Download Gain-of-Function Mutations in SCN5A Gene Lead to Type

Document related concepts

Cellular differentiation wikipedia , lookup

Cell encapsulation wikipedia , lookup

Organ-on-a-chip wikipedia , lookup

Amitosis wikipedia , lookup

List of types of proteins wikipedia , lookup

VLDL receptor wikipedia , lookup

Transcript
Cleveland State University
EngagedScholarship@CSU
ETD Archive
2012
Gain-of-Function Mutations in SCN5A Gene Lead
to Type-3 Long QT Syndrome
Fang Fang
Cleveland State University
How does access to this work benefit you? Let us know!
Follow this and additional works at: http://engagedscholarship.csuohio.edu/etdarchive
Part of the Chemistry Commons
Recommended Citation
Fang, Fang, "Gain-of-Function Mutations in SCN5A Gene Lead to Type-3 Long QT Syndrome" (2012). ETD Archive. Paper 94.
This Dissertation is brought to you for free and open access by EngagedScholarship@CSU. It has been accepted for inclusion in ETD Archive by an
authorized administrator of EngagedScholarship@CSU. For more information, please contact [email protected].
GAIN-OF-FUNCTION MUTATIONS IN SCN5A GENE LEAD TO TYPE-3 LONG QT
SYNDROME
FANG FANG
Bachelor’s of Science in Chemistry
Zhengzhou University
August 2004
Master’s of Science in Chemistry
Cleveland State University
December 2009
submitted in partial fulfillment of requirement for the degree
DOCTOR OF PHILOSOPHY IN CLINICAL AND BIOANALYTICAL CHEMISTRY
at the
CLEVELAND STATE UNIVERSITY
NOV 2012
@COPYRIGHT BY FANG FANG 2012
This dissertation has been approved for
the Department of Chemistry and for the College of Graduate Studies of
Cleveland State University
By
________________________________________________
Dr. Qing Wang, Department of Molecular Cardiology/CCF
Major Advisor
Date ________________
________________________________________________
Dr. Yan Xu, Department of Chemistry/CSU
Academic Advisor
Date ________________
________________________________________________
Dr. Xue-long Sun, Department of Chemistry/CSU
Advisory Committee Member
Date ________________
________________________________________________
Dr. Robert Wei, Department of Chemistry/CSU
Advisory Committee Member
Date ________________
________________________________________________
Dr. Yuping Wu, Department of Mathematics/CSU
External Examiner
Date ________________
ACKNOWLEGEMENTS
I would like to take this opportunity to express my sincere appreciation to everyone
who has helped me in the past six years. I would especially thank my research advisor, Dr.
Qing Wang, who gave me opportunities to study these interesting projects. During my
five years of study in Dr. Wang’s lab, not only did he have provided all the possible
experimental environment and let me finish my research study, but he also showed his
patience, encouragement and financial support.
Besides, I would like to thank Dr. Teng Zhang and Dr. Sandro Yong in our lab who
gave me a lot of advices and trained me to learn experiment techniques I used in my
research. I would like to thank Dr. Carlos Oberti and Xiaofen Wu, the experts in
Cardiology, to give me a lot of advice in Electrocardiogram (ECG) readings.
I also would like to thank all my committee members, Dr. Yan Xu, Dr. Robert Wei,
Dr. Xue-long Sun and Dr. Yuping Wu for their revision and suggestion of this
dissertation.
Finally, I would like to thank my husband, Dr. Xiaopeng Li and my parents.
Without their support, I cannot complete this dissertation. I would also like to thank my
son, Yifang Li, being such a wonderful boy and using his sweetest smile to support me.
GAIN-OF-FUNCTION MUTATIONS IN SCN5A GENE LEAD TO TYPE-3 LONG QT
SYNDROME
FANG FANG
ABSTRACT
Type-3 long QT syndrome, which is related to type 5 voltage-gated sodium channel
alpha subunit (SCN5A) mutation, has been identified since 1995. LQTS mutation in
SCN5A is a gain-of-function mutation producing late sodium current, INa,L. Brugada
mutation in SCN5A is a loss-of-function causing INa decrease. Whereas, the mechanism
for Dilated Cardiomyopathy mutations in SCN5A is still not fully understood.
N1325S is one of the first series of mutations identified for type-3 LQTS. Our lab
created a mouse model for LQTS by expressing SCN5A mutation N1325S in the mouse
hearts (TG-NS) and a matched experimental control line with overexpression of wildtype SCN5A (TG-WT). There are some interesting findings in TG-NS mice: (i)
Intracellular sodium (Na+) level is higher in TG-NS myocytes compared with TG-WT
myocytes. (ii) Ca2+ handling is abnormal in TG-NS myocytes, but not in TG-WT
myocytes. (iii) Apoptosis was also found in TG-NS mouse heart tissue, but not in TG-WT
hearts. These results provoke the hypothesis that gain-of-function mutation N1325S in
SCN5A leads to LQTS through abnormal cytosolic Ca2+ homeostasis.
Another LQTS mutation in SCN5A R1193Q was identified in 2004 and the
electrophysiological property is similar to other gain-of-function SCN5A mutations. The
transgenic mouse model for this mutation was also established and the surface
v
Electrocardiogram (ECG) results indicate longer corrected QT interval also present in
transgenic mice carrying R1193Q mutation. Besides, quinidine, an anti-arrhythmic
medication, can cause arrhythmic symptoms such as premature ventricular contraction
(PVC), premature atrial contraction (PAC) and atrioventricular (AV) block in R1193Q
transgenic mice.
In order to further study the relationship between abnormal Ca2+ handling and the
type of SCN5A mutation, either gain-of-function or loss-of-function, we have chosen
HL-1 cells, a cell line with indefinite passages in culture with all the adult cardiac
phenotypes. The similar abnormal Ca2+ handling was also identified in HL-1 cells
expressing N1325S mutation but not in those cells expressing wild-type SCN5A gene.
Since we hypothesized that the abnormal Ca2+ handling is caused by INa,L created by
gain-of-function mutation, either in HL-1 cells or in isolated TG-NS myocytes, I then use
INa,L blocker ranolazine, a clinical trial medication for LQT patients, to specifically block
INa,L. After the blockage of INa,L, the abnormal Ca2+ handling was rescued in both isolated
myocytes from TG-NS mice and HL-1 cells expressing N1325S mutation. Finally,
several different types of SCN5A mutations related to different types of heart diseases
were selected and the Ca2+ handling was tested in transfected HL-1 cells.
vi
TABLE OF CONTENTS
ABSTRACT …………………………………………………………....…………………v
LIST OF TABLES ………………………….…...………………….………….……….xiii
LIST OF FIGURES ……………………………..……………..…………………..…...xiv
LIST OF ABBREVIATIONS…………………………..……….………………....…….xv
CHAPTER
I. INTRODUCTION ………………………..……..…………………..…………...1
1.1 Voltage-Gated Sodium Channel Gene SCN5A ………..……..……….…1
1.1.1 The Family of Voltage-Gated Sodium Channels ……...……..…1
1.1.2 Type 5 Voltage-Gated Sodium Channel Nav1.5 and SCN5A
Gene ……………………………………………………….…...5
1.2 Cardiac Disease and Mutations in SCN5A Gene …..………..…….…….6
1.2.1 Long QT Syndrome …..…………………………..……….……7
1.2.2 Dilated Cardiomyopathy ………………………..…………..…9
1.2.3 Brugada Syndrome …………………………………..…....….....9
1.2.4 Progressive Cardiac Conduction Defect (PCCD)…………..….11
1.2.5 Sick Sinus Syndrome (SSS) …………...……………….…...…13
1.2.6 Atrial Fibrillation ………………………….…...…….………..14
1.2.7 Overlapping Syndromes …………………………………...….15
1.3 Cardiovascular System ………………..…………..……….…………..15
1.3.1 The Structure of the Cardiovascular System ..……….……..…15
vii
1.3.2 Electrical Conduction in the Heart and Electrocardiogram
(ECG) …………………………………………………………17
1.4 HL-1 Atrial Cells …………………………………………..……..…….17
1.4.1 Development of HL-1 Cardiomyocytes ……....……………….17
1.4.2 Studies of Cardiac Muscle Cell Structure and Function
Using HL-1 Cardiomyocytes ………………….………....…….18
1.5 Regulation of Nav1.5 Function in Cardiac Cells ……………...........…..19
1.5.1 Ankyrin Proteins ……………..………….…………...………..19
1.5.2 Caveolin-3 ……………..………………..………….....……….19
1.5.3 Nav1.5 β-Subunits ………………..……….……..………….…20
1.5.4 Protein Kinase A (PKA) …………..………………..….………21
1.5.5 Tyrosine Phosphorylation of Nav1.5 …………..…..……..…....22
1.5.6 Glycerol-3-Phospate Dehydrogenase Like Protein (GPD1-L)...23
1.5.7 Calmodulin (CaM) ………………..…………….......…………23
1.5.8 Ca2+/Calmodulin-Dependent Protein KinaseII (CaMKII)…...24
1.5.9 MOG1 …………………..………………………...……...……25
1.5.10 Fibroblast Growth Factor Homologous Factor 1B (FHF1B or
FGF12-1b) …………………………..…..……….………..….25
1.5.11 14-3-3η Protein ……………..……………….………...….….26
1.5.12 Ubiquitin-Protein Ligases of the Nedd4/Nedd4-Like Family...26
1.6 Anti-Arrhythmic Medications and Their Side Effects ……..…….….....27
viii
1.6.1 Quinidine …………..……………………………..…………....27
1.6.2 Mexiletine …………………..………………….……..……….28
1.6.3 Ranolazine …………………..……………………..……..……29
1.6.4 Other Anti-Arrhythmic drugs......................................................30
1.7 Cardiac Excitation-Contraction Coupling (ECG) and Ca2+ Handling in
Cardiac Cells ……………………………………………………..…...30
1.7.1 The Importance of Ca2+ in Cardiomyocytes …………........…..30
1.7.2 The Process of Excitation-Contraction Coupling …………......31
1.7.3 Cardiac Ion Channels and Cardiac Action Potential ..………....33
1.7.4 The Cellular Relaxation and the Removal of Cytosolic Ca2+.…33
1.7.5 Dysfunction of Na+/Ca2+ Exchanger Causes Ca2+ Overload and
Arrhythmias ……………………………………………..........33
1.7.6 Intracellular Ca2+ Transient and Excitation-Contraction
Coupling ……………………………………………………....36
II CARDIAC EXPRESSION OF SCN5A SNP R1193Q IN TRANSGENIC MICE
PRLONGS QT INTERVAL ………………………...….…………….……...…38
2.1 Abstract ………………………………..……………….….……..….…38
2.2 Introduction ……………………………..……………………………...39
2.3 Materials and Methods ……………………………..…………………..40
2.3.1 Preparation of Experimental Animals...…………................…..40
2.3.2 Western Blot Analysis …………………..…………...……..….42
ix
2.3.3 Semi-Quantitative Real-Time PCR (RT-PCR) Analysis…….....43
2.3.4 Telemetry ECG Recordings ………………..……….……..…..43
2.3.5 Echocardiographic Assessment of Cardiac Structure and
Function in Transgenic Mice ………………………….........….44
2.3.6 Data Analysis .............................................................................44
2.4 Results …………………..………………..………………………...…..45
2.4.1 Generation and Identification of Transgenic Mice with
Cardiac Specific Overexpression of Mutant Human SCN5A
Gene with SNP R1193Q..............................................................45
2.4.2 Development of Long-QT Syndrome in TG-RQ Mice………..47
2.4.3 Quinidine Prolonged QTc, and Increased the Frequencies of
PACs, PVCs, and Sinus Atrial Exit Block in TG-RQ Mice.........49
2.4.4 Echocardiographic Assessment of Cardiac Structure and
Function in the Mice …………….…………………….….....…54
2.5 Discussion ………………..………………………………..……..….....56
III LQTS MUTATION N1325S IN SCN5A CAUSES DISRUPTION OF
CYTOSOLIC CALCIUM HOMEOSTASIS IN MOUSE VENTRICULAR
MYOCYTES .……………………………………………………………….....59
3.1 Abstract ……………………………..…………..……………………...59
3.2 Introduction ………………………..…………………..…….……..…..60
3.3 Methods and Materials …………………………..………..………..…..62
x
3.3.1 Isolation of Mouse Ventricular Myocytes …………………..…62
3.3.2 Measurements of Ca2+ Transient Signals ……..……..…..…….64
3.4 Results ………………..……………………………….………........…..65
3.5 Discussion …………..…………………………………..…………...…69
IV LATE SODIUM CURRENT PRODUCED BY GAIN-OF-FUNCTION LQTS
MUTATIONS IN CARDIAC SODIUM CHANNEL LEADS TO ABNORMAL
CALCIUM HANDLING IN CARDIAC CELLS .…..………………………....72
4.1 Abstract …………………………………..……….…………….…..….72
4.2 Introduction ………………………………..…………..…………..…...73
4.3 Materials and Methods ……………………………..……….………….75
4.3.1 HL-1 Cardiac Cell Culture …………………………….............75
4.3.2 Na+ Channel Cloning and Site-Directed Mutagenesis …….......76
4.3.3 HL-1 Cell Transfection ….…………………………..….……..76
4.3.4 Intracellular Ca2+ Transient Measurements ….………...……...77
4.3.5 Data Analysis …………………………………………...……..77
4.4 Results ………………………………………………………..………...78
4.4.1 N1325S Mutant Sodium Channels Lead to Irregular
Ca2+ Handling in HL-1 Cells ……………………………..…..78
4.4.2 Ranolazine Can Reduce the SR Ca2+ Content and Strengthen
the Ability of Removing Extra Cytosolic Ca2+ in HL-1 Cells
Expressing N1325S Mutant SCN5A …………………...…..…81
xi
4.4.3 Mexiletine Can Reduce the SR Ca2+ Content in HL-1 Cells
Expressing N1325S Mutant SCN5A ……………………….....85
4.4.4 Classification of Different SCN5A Mutations …………….…...87
4.4.5 Ranolazine Effect in Various SCN5A Mutations……….........…89
4.5 Discussion ………………………………..…………………..……..….91
BIBLIOGRAPHY ..........………………………………………………………..….……94
xii
LIST OF TABLES
Table
I.
page
Different Types of Voltage-Gated Sodium Channel Alpha Subunit …………….…...3
II. Overlapping Syndrome of Cardiac Sodium Channelpathy ………………….….…..16
III. Ionic Currents and Different Phases of AP ………………………………….……...35
IV. Effects of Quinidine Treatment on PACs, PVCs and SAEB in TG-RQ Mice and
Control NTG Mice......................................................................................................53
V. Echocardiographic Assessment of Cardiac Structure and Function in the Mice…....55
xiii
LIST OF FIGURES
Figure
Page
1. Schematic Diagram of Cardiac Excitation-Contraction Coupling (ECG) Events
in Ventricular Myocytes ………………………………………………....………..…32
2. A schematic Diagram Showing that Different Ionic Currents Contribute to
Different Phases of a Cardiac Action Potential ………………………….………..…34
3. Generation and Genotyping of TG-RQ Mice ………………………………….……46
4. ECG Recordings of Non-Transgenic(NTG), Transgenic Wild-Type (TG-WT)
and Transgenic R1193Q (TG-RQ) Mice ……………………………….…...…….…48
5. Quinidine Treatment Prolonged QTc in Both NTG and TG-RQ Mice......…….….…50
6. Quinidine Treatment Induced Increased PVCs, PACs and SAEB in TG-RQ mice.....52
7. Ca2+ Transients Signals in Isolated TG-NSL3 Mice Myocytes …………………..…67
8. Ca2+ Transients Signals in Isolated TG-NSL12 Mice Myocytes ……………….…...68
9. Ca2+ Transient Signals in Transfected HL-1 Cells …………………………….….…80
10. Effect of Ranolazine Treatment in HL-1 Cells Expressing the N1325S Mutation ….84
11. Effect of Mexiletine treatments in HL-1 Cells Expressing the N1325S Mutation.….86
12. Effects of Different SCN5A Mutations on Ca2+ Transients …………………………88
13. Effect of Ranolazine in Various SCN5A Mutations ……………………………..…90
xiv
LIST OF ABBREVIATIONS
AF
Atrial Fibrillation
AP
Action Potential
AV node
Atrioventricule node
BDM
2,3-Butanedione Monoxime
BrS
Brugada Syndrome
CaM
Calmodulin
CaMKII
Ca2+/Calmodulin-Dependent Protein Kinase II
CCD
Cardiac Conduction Disturbance
CICR
Ca2+ Induced Ca2+ Release
CNS
Central Nerve System
DAD
Delayed Afterdepolarization
DCM
Dilated Cardiomyopathy
DRG
Dorsal Root Ganglion
EAD
Early Afterdepolarization
ECC
Excitation-Contraction Coupling
ECG
Electrocardiogram
EGF
Epidermal Growth Factor
EGFR
Epidermal Growth Factor Receptor
ER
Endoplasmic Reticulum
xv
FHF1B
Fibroblast Growth Factor Homologous Factor 1B
GPD1-L
Glycerol-3-Phosphate Dehydrogenase Like Protein
IQ
Isoleucine-Glutamine
IVF
Idiopathic Ventricular Fibrillation
LQTS
Long QT Syndrome
MHC
Myosin Heavy Chain
MOG1
Multicopy Suppressor of gsp1 Mutants
NCX
Na+/Ca2+ Exchanger
PACs
Premature Atria Contractions
PCCD
Progressive Cardiac Conduction Defect
PKA
Protein Kinase A
PKC
Protein Kinase C
PVCs
Premature Ventricle Contractions
PVDF
Polyvinylidene Difluoride
RT-PCR
Real-Time Polymerase Chain Reaction
SA node
Sinoatrial node
SERCA
Sarcolemmal Ca2+-ATPase
SIDS
Sudden Infant Death Syndrome
SND
Sinus Node Dysfunction
SNP
Single Nucleotide Polymorphism
xvi
SR
Sarcoplasmic Reticulum
SSS
Sick Sinus Node Syndrome
TN-C
Troponin Complex
TTX
Tetrodotoxin
VF
Ventricular Fibrillation
VT
Ventricular Tachycardias
xvii
CHAPTER I
INTRODUCTION
1.1 Voltage-Gated Sodium Channel Gene SCN5A
1.1.1
The Family of Voltage-Gated Sodium Channels
Voltage-gated sodium channels are large membrane proteins which produce
depolarization and initiate action potentials in neurons, cardiomyocytes and skeleton
muscle cells (Andavan and Lemmens-Gruber 2011). The sodium channels are composed
of a large pore forming α subunit and one or several small β subunits (Abriel 2007). To
date, there are at least twelve different α subunit isoforms and four β subunit isoforms
that have been cloned for voltage-gated sodium channels in mammals.
Voltage-gated sodium channel α subunit contains four homologous domains (DI to
DIV), and each domain is composed of six transmembrane segments (S1 to S6)
(Abriel 2007). The fourth transmembrane segment (S4) carries positively charged amino
acids and plays an important role as a voltage sensor (Abriel and Kass 2005). The pore is
1
formed by the extracellular loop between S5 and S6 (Schroeter et al 2010). The
site-directed mutation study in the sodium channel α subunit indicates that the activation
of the channel is regulated by S6 (Hanlon and Wallace 2002). The fast inactivation
involves the linker between DIII and DIV; whereas the slow inactivation is modulated by
the conformational change of the extracellular region of the channel (Hanlon and Wallace
2002, Ragsdale et al 1994).
Twelve isoforms of voltage-gated sodium channel α subunits are expressed in
different tissues (Table I). The dysfunction of sodium channels causes a variety of human
diseases. SCN1A, SCN2A and SCN3A are mainly expressed in neurons, and the mutations
in these three sodium channels can cause seizures and epilepsy, especially in infants
(Baulac et al 1999, Bender et al 2012, Kearney et al 2001, Saitoh et al 2012, Wallace et al
2001). Mutations in skeleton muscle sodium channel gene SCN4A cause hyperkalemic
periodic paralysis and paramyotonia congenital (Meyer-Kleine et al 1994, Plassart et al
1994). Mutations in cardiac sodium channel gene SCN5A produce persistent late sodium
current (I Na,L ) and are the main cause of long-QT syndrome (LQTS), ventricular
fibrillation (VF) and Brugada syndrome (Baroudi et al 2000, Bezzina et al 1999, Chen et
al 2002, Deschênes et al 2000, Tian et al 2004, Tian et al 2007, Wan et al 2001, Wang et
al 1995a, Wang et al 2004, Zhang et al 2011). In 1990, SCN6A was identified in human
heart and uterus (George et al 1994). Later, SCN7A was assigned although the authors
symbolized this gene as SCN6A, and the localization of SCN7A (SCN6A) was obtained
using chromosome microdissection-PCR method (George et al 1994, Han et al 1991).
The single nucleotide polymorphism in human SCN7A gene was found to be associated
with essential hypertension in the Chinese population (Zhang et al 2003). SCN8A
2
Table I. Different Types of Voltage-Gated Sodium Channel Alpha Subunit
Gene
Chromosome
Localization
Tissue
SCN1A
2q23-q24
Brain
SCN2A
2q22-q23
Brain
SCN3A
2q24-q31
Brain and
dorsal root
ganglion (DRG)
SCN4A
17q23.1-q25.3
SCN5A
Mutation and
Disease
Epilepsy and
Dravet’s syndrome
Seizures
References
(Saitoh et al 2012)
(Bender et al 2012)
(Wallace et al 2001)
(Kearney et al 2001)
Epilepsy
(Baulac et al 1999)
Skeletal muscle
Paramyotonia and
congenita
(Plassart et al 1994)
(Meyer-Kleine et al
1994)
3p21
Heart
LQTS
(Wang et al 1995b)
(Bezzina et al 1999)
SCN6A
2q21-q23
Uterus and heart
Essential
hypertension
(George et al 1994)
(Zhang et al 2003)
SCN7A
2q21-q23
Glial cells and
DRG
N/A
(Potts et al 1993)
SCN8A
12q13
Brain, DRG and
spinal cord
Cerebellar atrophy,
ataxia and mental
retardation
(Trudeau et al 2006)
SCN9A
2q24
DRG, brain and
spinal cord
Insensitive to pain
(Nilsen et al 2009)
SCN10A
3p22-p24
DRG and heart
Longer PR on ECG
(Chambers et al
2010)
SCN11A
3p21-p24
DRG
N/A
(Jeong et al 2000)
SCN12A
3p23-p21.3
Central nerve
system (CNS)
and DRG
N/A
(Jeong et al 2000)
3
was found in fetal brains and the protein contains 1,980 amino acids. Interestingly, human
SCN8A shares 98.5% sequences with the same mouse gene (Plummer et al 1998). SCN8A
gene was first studied in mutant mice. Mutations in SCN8A in mice caused dysfunction in
motor units such as “motor endplate disease” and mutant jointing (Burgess et al 1995,
Kohrman et al 1996). A patient with a heterozygous two base pairs deletion mutation in
exon 24 of SCN8A was found to have cerebellar atrophy, ataxia, and mental retardation
(Trudeau et al 2006). Higher resistance to seizures in two mutant Scn8a mouse lines,
Scn8a (med) and Scn8a (med-jo) suggests that mutant Scn8a can reduce the neural
excitability (Martin et al 2007). Mutations in SCN9A were found in patients insensitive to
pain, which is termed as “channelopathy -associated insensitivity to pain” (Nilsen et al
2009). Genome wide association studies of SCN10A indicated that SNPs in this gene can
lead to longer PR interval on electrocardiograms (ECG) (Chambers et al 2010). Scn11a
was first identified in the mouse and expressed in the small sensory neurons of dorsal
ganglia (DRG) and trigeminal ganglia (Dib-Hajj et al 1998, Dib-Hajj et al 1999). Later,
SCN11A, which was called SCN12A, was identified in the human brain and found to be
expressed in the central nervous system (CNS) especially in the gray matter of the
hippocampus, cerebellum and olfactory bulb (Jeong et al 2000).
Although different types of sodium channels are expressed in different species and
tissues, they are highly conserved evolutionarily. Cardiac sodium channel gene, SCN5A,
is especially highly homologous to SCN8A and SCN9A which are primarily expressed in
neurons (Kerr et al 2004, Plummer and Meisler 1999). SCN6A and SCN7A are actually
orthologs in the human and mouse (George et al 1990, George et al 1994, Han et al 1991).
SCN1A, SCN2A and SCN3A are form a gene cluster on chromosome 2q24 (Mulley et al
4
2005). SCN5A, SCN10A and SCN11A are evolutionally more closely related because they
are tetrodotoxin-sensitive, and all located on chromosome 3p21 (Plummer and Meisler
1999). The sequence of SCN8A is also highly similar to different domains in SCN2A,
SCN4A and SCN5A. For example, the second domain in SCN8A is 93%, 89% and 80%
homologous to SCN2A, SCN4A and SCN5A, respectively. However, the cytoplasmic loop
between Domain I and Domain II in SCN8A is less identical to the other three genes
(Plummer and Meisler 1999).
The β subunits for voltage-gated sodium channels are glycoproteins which are not
only the accessory part for alpha subunits, but also modulate Na+ current (INa ) (Patino
and Isom 2010). In addition, they function in cell aggregation, migration, invasion and
neurite outgrowth when working as cell adhesion molecules (CAM) (Schroeter et al
2010). There are at least four types of voltage-gated sodium channel β subunits referred
to as β1-β4, which are encoded by SCN1B-SCN4B, respectively (Patino and Isom 2010).
1.1.2 Type 5 Voltage-Gated Sodium Channel Na v 1.5 and SCN5A Gene
Type 5 voltage-gated sodium channel, Na v 1.5, is an integrated membrane protein
which mediates the rapid upstroke of the action potential (AP) in cardiac cells. Na v 1.5 is
the pore forming subunit of human cardiac sodium channel. The entrance of Na+ through
Na v 1.5 into a cardiac cell initiates the action potential in both ventricular and atrial
cardiac cells (Schroeter et al 2010). Na v 1.5 is encoded by the SCN5A gene which is
located on chromosome 3p21 and is composed of 22 introns and 28 exons (George et al
1995). Na v 1.5 is a large protein containing 2,016 amino acids with an estimated
molecular weight of 227 kDa (Wang et al 1996b). Northern blot analysis showed that this
5
gene was mainly expressed in the heart (Gellens et al 1992).
Currently, there are several alternatively spliced Na v 1.5 variants in different
species and tissues. Na v 1.5a was first observed in rat brain, and was the first cardiac
sodium channel splice variant. Later, Na v 1.5a was found in diverse tissues in rodents, but
was not expressed in human heart (Schroeter et al 2010). Na v 1.5c and Na v 1.5d are two
splice variants in human hearts. The expression of Na v 1.5c is more abundant than
Na v 1.5d. In Na v 1.5c, an extra glutamine residue is introduced at codon position 1077 on
the 5’-end of exon 18. The transcriptional ratio of Na v 1.5c: Na v 1.5 is around 1:2, and
there is no obvious functional differences in Na v 1.5c compared with Na v 1.5 (Makielski
et al 2003). In Na v 1.5d, a 120 bp fragment is deleted from exon 17, and as a result, 44
amino acids are missing on the intracellular loop II between the second segment DII and
the third segment DIII. Electrophysiological study has demonstrated that the channel
kinetics is significantly changed in Na v 1.5d (Camacho et al 2006, Gellens et al 1992).
1.2 Cardiac Diseases and Mutations in SCN5A Gene
In 1995, the first mutation in the SCN5A gene was identified and linked to an
inherited cardiac arrhythmia, long QT syndrome (LQTS) (Wang et al 1995b). Since then,
various mutations in this gene have been found and associated with numerous arrhythmic
syndromes, such as Brugada syndrome (BrS), progressive cardiac conduction defect
(PCCD), sick sinus node syndrome (SSS), atrial fibrillation (AF) and dilation
cardiomyopathy (DCM) (Beggs 1997, Chen et al 1998, Lei et al 2008, Makiyama et al
2008, Tan et al 2001).
6
1.2.1
Long QT Syndrome
LQTS is a cardiac disorder characterized by the prolonged QT interval on the
surface electrocardiogram (ECG), syncope and sudden death as a result of a polymorphic
ventricular tachycardia, torsade de pointes (Wang et al 2004). Both genetic and acquired
or environmental factors contribute to the pathogenesis of LQTS. To date, LQTS is
known to be caused by mutations in at least thirteen genes, including the cardiac
potassium channel genes KvLQT1 or KCNQ1 (11p15.5, LQT1) (Neyroud et al 1997,
Splawski et al 1997a, Wang et al 1996a), HERG or KCNh2 (7q35-36, LQT2) (Curran et
al 1995), the cardiac sodium channel gene SCN5A (3p21-24, LQT3) (Wang et al 1995a),
the non-ion channel ankyrin-B gene encoding a protein that links ion channels to the
cytoskeleton (4q25-27, LQT4) (Mohler et al 2003), cardiac potassium channel gene
KCNE1 (21q22, LQT5) (Schulze-Bahr et al 2003, Splawski et al 1997b), KCNE2 (21q22,
LQT6) (Abbott et al 1999), KCNJ2 (17q24.3, LQT7) (Tristani-Firouzi et al 2002), L-type
Ca2+ channel CACNA1C (12p13.3, LQT8) (Navedo et al 2010), scaffolding protein
caveolin-3 CAV3 (3p25, LQT9) (Balijepalli and Kamp 2008), the cardiac sodium channel
gene SCN4B (11q22.3, LQT10) (Vincent 2003), A-kinase-anchoring protein-9 (AKAP9)
(7q21.2, LQT11) (Chen et al 2007), alpha-1 syntrophin gene (SNTA1) (20q11.21, LQT12)
(Ueda et al 2008), and potassium channel gene KCNJ5 (11q24, LQT13) (Yang et al
2010).
In 1995, Wang et al. reported that mutations in the cardiac sodium channel gene
SCN5A caused type 3 LQTS. To date, there are more than 45 mutations in SCN5A that
have been identified to be causative to LQTS. These mutations span the entire sodium
channel from the 9th amino acid (G9V) to the last amino acid (P2006A) (Wang et al 2007).
7
The first gain-of-function mutation ΔKPQ in SCN5A produced late Na+ current (I NaL ) and
was first identified in two unrelated LQTS families. ΔKPQ is a three amino acids deletion
at position 1505-1507 locating at the intracellular loop between DIII and DIV (Wang et al
1995b). Because of the persistent late Na+ current (I NaL ), which is cause by the
fluctuation of this mutant channel between normal and non-inactivating modes, the
cardiac action potential is prolonged, which provides a potential mechanism for this type
of congenital LQTS (Bennett et al 1995). N1325S is another gain-of-function mutation in
SCN5A gene causing type 3 LQTS. A transgenic mouse model which expressed mutant
human N1325S SCN5A gene was established and showed prolonged QT interval on ECG,
spontaneous ventricular tachyarrhythmia (VT) and ventricular fibrillation (VF) (Tian et al
2004).
In 2004, a missense mutation in SCN5A, R1193Q, was found in one of seven
patients with drug-induced acquired LQTS. This mutation can destabilize inactivation
gating and generate a persistent, non-inactivating current in both human kidney
(HEK-293) cells and Xenopus oocytes. The biophysical defects of R1193Q are almost
identical to two well-characterized SCN5A mutations, N1325S and R1644H
(Wattanasirichaigoon et al 1999), which can cause type 3 LQTS. One group reported that
SCN5A R1193Q is a polymorphism in a Chinese family that is associated with
progressive cardiac conduction defects and long QT syndrome (Sun et al 2008). Another
group reported that SCN5A R1193Q is a common polymorphism in general Chinese
population. They pointed out that about 12% (11/94) of the subjects carried the R1193Q
mutation. They analyzed clinical data from 9 of the 11 patients with the mutation.
Although none of the carriers had Brugada syndrome, one had prolonged QTc and
8
another had borderline QTc appeared to be homozygous for R1193Q (Hwang et al 2005).
1.2.2
Dilated Cardiomyopathy
Patients with idiopathic dilated cardiomyopathy (DCM) usually have clinical
features of an increased ventricular size and reduced ventricular systolic function (Beggs
1997). Most mutations causing DCM are thought to have either structural defects or
dysfunction in electrical excitation-contraction in cardiac cells. The structural defects
caused by DCM mutations usually result in contractile filament disarray and cell death,
and further lead to fibrosis which is the characteristics of DCM (Asahi et al 2003, Chen
and Chien 1999, Kane et al 2005, Schmitt et al 2003).
In 1996, SCN5A was first related to DCM after the study of a family with DCM
associated with sinus node dysfunction, supraventricular tachyarrhythmias, conduction
delay, and stroke (Olson and Keating 1996). Although SCN5A mutations related to DCM
are not as common as to other cardiac diseases, the high percentage of prevalence of
familial DCM indicates that gene defects play an important role in this cardiac disorder.
To date, around ten mutations in SCN5A that cause DCM have been reported. However,
the molecular mechanism by which these mutations can cause DCM is still not clear
(Ruan et al 2009). It was also suggested that some mutations in SCN5A such as N1325S
can cause both LQTS and DCM, and this result has already been confirmed in transgenic
mouse model expressing N1325S mutant hSCN5A (Yong et al 2007, Zhang et al 2011).
1.2.3
Brugada Syndrome
Brugada syndrome (BrS) is a cardiac disease that is characterized by elevated
9
ST-segment in right precordial leads (V1-V3) on ECG, and it is a common form of
idiopathic ventricular fibrillation (IVF). Patients with BrS will have increased risk in
cardiac arrhythmias and sudden death caused by episodes of rapid polymorphic
ventricular tachycardias (VT) or ventricular fibrillation (VF) (Brugada and Brugada
1992). Although some patients with BrS have no symptoms on their ECG (IVF with
normal ECG), their symptoms can be accentuated by the administration of some
medications such as sodium channel blocker flecainide (Brugada et al 2002).
Cardiac sodium channel SCN5A related to type 1 BrS. In 1998, the first series of
SCN5A mutations were identified in BrS patients. T1620M is a missense mutation on the
extracellular loop between S3 and S4 on DIV (Chen et al 1998). Nowadays, more than 90
mutations in SCN5A are identified to be linked to BrS (Ruan et al 2009). Opposite to
gain-of-function LQTS mutations in SCN5A, BrS mutations are loss of function
mutations. Chen et al. stated that missense mutation T1620M locating on the extracellular
loop between S3 and S4 on DIV exhibited a significantly faster recovery compared with
wild-type channels. Later, the trafficking of sodium channel α subunit is also considered
as another potential mechanism for BrS. In 2000, a missense mutation R1432G was
reported to reduce the membrane sodium channel expression and sodium current density
(Deschênes et al 2000).
Besides the SCN5A gene, there are six other genes that are related to BrS. GPD1L
(3p21-23) encoding a protein that interacts with sodium channel α subunit can cause type
2 BrS (BrS2). CACNA1C (12p13.3) encoding calcium channel α subunit (Ca v 1.2) can
cause BrS3. CACNB2 encoding calcium channel β subunit (Ca v β2) causes BrS4. SCN1B
(19q13.1) encoding one of the sodium channel β subunit (Na v β) causes BrS5. KCNE3
10
(11q13-q14) encoding β subunit of I Ks /I to (MiRP2) causes BrS6. SCN3B (11q23.3)
encoding type 3 β subunit of sodium channel (Na v β3) causes BrS7 (Hedley et al 2009).
It is estimated that the prevalence of BrS is about 5:10,000 (Blangy et al 2005).
BrS is also thought to be responsible for at least 4% of all sudden deaths and 20% of all
sudden deaths in patients without any heart structural abnormality (Antzelevitch et al
2007). As a result, the study of BrS mechanism is very essential. Since the BrS mutations
in SCN5A are caused by the decreased membrane channel expression, rescuing the
trafficking defect might be a potential treatment for BrS (Bezzina and Tan 2002, Valdivia
et al 2002).
1.2.4
Progressive Cardiac Conduction Defect (PCCD)
Progressive cardiac conduction defect (PCCD), also known as Lev-Lenegre disease,
is one of the most common cardiac conduction disturbances (CCD) and characterized by
progressive alteration of cardiac conduction through the blockage of left and right HisPurkinje bundles and QRS complex widening (Schott et al 1999). Patients with PCCD
ultimately develop chronic atrio-ventricular block due to the abnormal His-Purkinje
system. As a result, pacemaker implantation is the most common treatment for patients
with PCCD (Royer et al 2005).
Cardiac conduction disturbance has been linked to the SCN5A gene and two other
chromosomal locations (19q13.2-13.3 and 16q23-24) where no genes have been
identified yet (Bezzina et al 2003, Brink et al 1995, de Meeus et al 1995, Schott et al
1999, Tan et al 2001). In 1999, Schott first reported two mutations in two families with
PCCD. In a large French family, a mutation of T→C substitution at position +2 of the
11
splicing donor site of intron 22 was found and expected to lead to in-frame skipping of
exon 22, producing an impaired protein without the voltage-sensor S4 in DIII. The young
individuals with this mutation did not show severe symptom. However, with aging and
the increases of fibrosis, patients will develop the PCCD phenotype. In another small
Dutch family, a deletion of a single nucleotide G at position 5280 was detected. The
mutation results in a frameshift which produces a premature stop codon. Individuals with
this mutation presented with a congenital conduction defect from birth, indicating that the
effect of this premature stop codon is immediate (Schott et al 1999). Later, a missense
mutation of G514C was identified in a 3-year old girl who experienced episodes of
fainting during a febrile disease. Her ECG revealed typical cardiac conduction defect
with slow conduction such as broad P waves, prolonged PR intervals and a wide QRS
complex. The electrophysiological study in TSA201 cells expressing the G514C mutation
showed an opposite effect on sodium current (INa ) during activation and inactivation. The
more rapidly decay of G514C channel reduces the I Na accessible for cardiac impulse
conduction. The positive shift (+7mV) of the normalized current-voltage (IV) curve
suggested the destabilized close-state inactivation, which would increase I Na (Tan et al
2001).
There are some PCCD mutations in SCN5A that can also cause Brugada syndrome,
which was described as cardiac sodium channel overlap syndromes (Remme et al 2008).
For example, a missense mutation G1406R was identified in a 45-member family. This
mutation can either cause Brugada syndrome or isolated cardiac conduction defect
(Kyndt et al 2001).
12
1.2.5
Sick Sinus Syndrome (SSS)
Sick sinus syndrome (SSS), also known as sinus node dysfunction (SND), is
characterized by variable manifestations that include inappropriate sinus bradycardia,
sinus arrest, atrial standstill, chronotropic incompetence and tachycardia bradycardia
syndrome. Patients with SSS always have symptoms such as dizziness or syncope
(Benson et al 2003, Dobrzynski et al 2007)
Besides atria and ventricles, SCN5A is also expressed in some pacemaker cells in
sinoatrial (SA) node. Although the function of I Na is still not very clear, some evidence
suggests that Na v 1.5 defects are the possible reason for some SA node dysfunction. To
date, at least 14 mutations in SCN5A associated with SSS have been recognized (Lei et al
2008, Ruan et al 2009). In 2003, six loss-of-function SCN5A mutations were identified in
five SSS patients. Two of these six mutations are completely nonfunctional (G1408R and
R1623X) and the other four are partially dysfunctional (T220I, P1298L, delF1617 and
R1632H) (Benson et al 2003). The study of SCN5A+/- knockout mice, which develop
symptoms such as sinus bradycardia, slowed SA conduction and SA exit block, indicates
that the loss-of-function of Na v 1.5 is the mechanism for SSS (Papadatos et al 2002).
There are also some gain-of-function mutations in SCN5A that can cause mixed
symptoms including SSS, LQTS, BrS and DCM. In 2003, a gain-of-function mutation
1795insD was found to produce both persistent sodium current and a negative shift in
inactivation. It is concluded that the effects of this persistent sodium current and the
negative shift reduce the sinus rate, thus explaining the mechanism for QT-prolongation
and sinus bradycardia and sinus pauses (Veldkamp et al 2003). Thus, both loss-offunction and gain-of-function mutations can cause the dysfunction of sinus node.
13
1.2.6
Atrial Fibrillation
Atrial fibrillation (AF) is the most common cardiac disorder that is characterized
by rapid and irregular activation of the atria. About 15% of all strokes are caused by AF.
One third of patients with strokes are over 65 years of age, thus as the population ages,
AF is responsible for the potential economic cost for the public health care system
(Fatkin et al 2007, Matsusue et al 2012, Wang et al 2003).
At least 11 genes are related to AF and they include both ion channel genes and
non-ion channel genes. Mutations in both potassium channels and sodium channels can
cause AF. The ion channel genes are KCNQ1 (α-subunit of I Ks channel), KCNE2
(β-subunit of I Ks channel), KCNE5 (β-subunit of I Ks channel), KCNJ2 (Kir2.1 channel),
KCNA5 (K v 1.5 channel), SCN5A (sodium channel α-subunit), SCN1B (sodium channel
β-subunit), SCN2B (sodium channel β-subunit), and SCN3B (Chen et al 2003, Das et al
2009, Hong et al 2005, Makiyama et al 2008, Olson et al 2005, Olson et al 2006, Otway
et al 2007, Ravn et al 2008, Wang et al 2010, Watanabe et al 2009, Yang et al 2004). The
non-ion channel genes are NUP155 (nucleoporin), GJA5 (connexin-40) and NPPA (atrial
natriuretic peptide (mANP)) (Gollob et al 2006, Hodgson-Zingman et al 2008, Oberti et
al 2004, Zhang et al 2008).
In 2008, a SCN5A mutation M1875T was found in a Japanese family with only AF.
This family has no symptoms of structural heart disease or ventricular arrhythmias. The
electrophysiological study showed that M1875T is a gain-of-function mutation without
producing persistent inward Na+ currents. The characteristic of this mutation is obviously
different from the LQT-type gain-of-function mutations which produce persistent late
sodium current (Makiyama et al 2008).
14
1.2.7
Overlapping Syndromes
The effects of mutations in SCN5A are usually complicated. As a result, patients
carrying certain mutations will display not only one type of symptom, but a mixed
syndromes. Numerous SCN5A mutations lead to mixed phenotypes, known as a
overlapping syndrome of cardiac sodium channelopathy (Remme et al 2008).
In 1999, Bezzina et al. first revealed that a single SCN5A mutation could cause
multiple cardiac disorders within the same family. They found out that the family
carrying 1795insD mutation in SCN5A presenting ECG features of sinus node
dysfunction, bradycardia, BrS and LQTS (Bezzina et al 1999). Later, another deletion
mutation ΔK1500 was reported to cause LQTS, BrS and conduction disease (Grant et al
2002). To date, more than 30 mutations in SCN5A can cause different combination of
cardiac diseases. Most of these combinations are cardiac conduction dysfunction (CCD)
and Brugada syndrome (BrS) (Table II).
1.3 Cardiovascular System
1.3.1
The Structure of the Cardiovascular System
The cardiovascular system is composed of the heart and blood vessels, which
supply oxygen and nutrition throughout the body. The heart is a muscular organ that
propels blood while contracting. There are four chambers in the human heart: left and
right atria and ventricles. There are two types of blood circulation systems in the human
body. Pulmonary circulation carries oxygen-deprived blood from the right ventricle
through the pulmonary artery to lungs, and then returns it back to the left atrium with
oxygenated blood through the pulmonary vein. Systematic circulation carries oxygenated
15
Table II. Overlapping Syndrome of Cardiac Sodium Channelpathy
SCN5A
Mutation
Clinical Phenotype
CCD
SSS
E161K
+
+
+
(Smits et al 2005)
T187I
+
+
(Makiyama et al 2005)
P336L
+
+
(Cordeiro et al 2006)
D365N
+
+
(Makiyama et al 2005)
R367H
+
+
(Takehara et al 2004)
R376H
+
+
(Rossenbacker et al 2004)
N406S
+
+
(Itoh et al 2005)
G752R
+
+
(Potet et al 2003)
F861fs951X
+
+
(Schulze-Bahr et al 2003)
E867X
+
+
(Schulze-Bahr et al 2003)
+
+
(Splawski et al 2000)
W1191X
+
+
E1225K
+
+
D1275N
P1332L
(Shin et al 2007)
+
+
G1319V
Other
References
BrS
D1114N
LQTS
+
+
(Schulze-Bahr et al 2003)
AF/DCM
(McNair et al 2004)
+
(Casini et al 2007)
+
(Ruan et al 2007)
I1350T
+
+
(Juang et al 2003)
G1406R
+
+
(Kyndt et al 2001)
G1408R
ΔK1479
ΔK1500
+
ΔKPQ
+
ΔF1617
+
+
+
(Kapplinger et al 2010)
+
+
+
(Schulze-Bahr et al 2003)
+
+
(Grant et al 2002)
+
(Wang et al 1995b)
+
+
(Chen et al 2005)
R1623X
+
+
(Makiyama et al 2005)
R1632H
+
+
(Benson et al 2003)
I1660V
+
+
(Cordeiro et al 2006)
S1710L
+
+
(Akai et al 2000)
V1763M
+
+
(Chang et al 2004)
M1766L
+
+
(Valdivia et al 2002)
V1777M
+
+
(Lupoglazoff et al 2002)
1795insD
+
+
+
(Bezzina et al 1999)
+
+
(Schulze-Bahr et al 2003)
+
(Shim et al 2005)
S1812X
P2005A
+
16
blood from the left ventricle through the aorta to the rest of the body, and then returns it
back to the right atria with oxygen-deprived blood (Luca et al 2009). The cardiac muscle
contraction is caused by the electrical signal turned by excitation-contraction coupling in
cardiac muscle cells.
1.3.2 Electrical Conduction in the Heart and Electrocardiogram (ECG)
The electrical signal arisen by cardiac action potentials starts at the sinoatrial
node (SAN), which is located in the upper side of the right atrium. SAN sends the
impulse to the right atrium, to the left atrium through Bachmann’s bundle, and then to the
atrioventricle (AV) node. The electrical signal of left and right atria generates the P wave
on the electrocardiogram (ECG). The AV node plays an important role in delaying the
ventricle contraction. Without the delay of AV node, ventricles and atria would beat at the
same time. Two branches of His bundles in AV node, left and right branches, activate two
ventricles, respectively and generate the PR segment on the ECG. The His bundles then
branch to produce a bunch of Purkinje fibers, which spread the electrical impulse in
ventricles, and form the QRS complex on the ECG (Klabunde 2012).
1.4 HL-1 Atrial Cells
1.4.1
Development of HL-1 Cardiomyocytes
In 1998, William C Claycomb successfully derived a cardiac muscle cell line,
named HL-1, from the AT-1 mouse atrial cardiomyocyte tumor lineage. HL-1 cells can be
passed infinitely, and still maintain the phenotype of cardiomyocytes. With specific
culture media, HL-1 cells can be cultured beyond 30 passages and still remain beating
17
and contain no other cell types. With immunochemistry tests, HL-1 cells were confirmed
to have perinuclear ANF-containing granules, cardiac-specific myosin and musclespecific desmin intermediate filaments. These immunochemistry observations confirmed
that HL-1 cells possess cardiac morphological properties. The gene expression analysis
using RT-PCR (reverse transcription PCR) identified the presence of adult isoform of
myosin (α-MHC) and contraction proteins such as α-cardiac actin, ANF and connexin43
(CX43). The gene expression results revealed that HL-1 cells retained biochemical
characteristics of cardiac cells. Finally, Claycomb reported that HL-1 cells showed
spontaneous action potentials and repolarizing K+ currents using whole cell patch-clamp
technique (Claycomb et al 1998).
1.4.2
Studies of Cardiac Muscle Cell Structure and Function Using HL-1
Cardiomyocytes
Before the development of HL-1 cardiomyocytes, isolated embryonic or neonatal
rat and mouse cardiomyocytes were most widely used for cardiac studies. However,
embryonic or neonatal cardiomyocytes lack the adult myocyte phenotypes. Some labs are
also using isolated adult cardiomyocytes, but these adult cardiomyocytes cannot be
cultured, and they are difficult to be transfected. The technique of adult cardiomyocytes
isolation also highly depends on the quality of enzymes. The invention of HL-1 cells, the
only cardiac cell line that can be continuously divided and still maintain the adult cardiac
phenotypes in culture, provides a model for the study of cardiac diseases (White et al
2004).
HL-1 cardiomyocytes were first used to study the change of gene expression in
18
cardiac cells under pathophysiological stress such as hypoxia, hyperglycemia, and
hyperinsulinemia (Collier et al 2002, Nguyen and Claycomb 1999). HL-1 cells were also
used to study cellular signaling pathways and different types of receptors in cardiomyocytes.
1.5 Regulation of Na v 1.5 Function in Cardiac Cells
1.5.1
Ankyrin Proteins
There are three ankyrin proteins encoded by ANK1, ANK2 and ANK3 genes.
Ankyrin-B (encoded by ANK2) and ankyrin-G (encoded by ANK3) are expressed in the
heart (Cunha and Mohler 2006). Although in ankyrin-B knockout mice, late opening of
Na v 1.5 was observed in isolated neonatal myocytes, there is no evidence showing that
ankyrin-B directly interacts with Na v 1.5 (Chauhan et al 2000). However, ankyrin-G was
found to directly interact with Na v 1.5 on the loop between DII and DIII on a 9 amino
acid conserved motif (Lemaillet et al 2003). In 2004, a BrS mutation in Na v 1.5, E1053K,
was reported by Mohler et al. This mutation is in the ankyrin-G binding motif and
disrupts the interaction between ankyrin-G and Na v 1.5. As a result, Na v 1.5 cannot be
transferred to plasma membrane, and tracked in the cytoplasm. These findings suggest
that ankyrin-G functions as a helper for Na v 1.5 trafficking and sorting (Mohler et al
2004).
1.5.2
Caveolin-3
The cell membrane is composed of lipids and proteins. According to Singer and
Nicholson’s fluid-mosaic model about the cell membrane, proteins can float in the
19
phospholipid bilayers freely (Brown and London 1998). The less fluid liquid-order phase
(Lo) is created by phospholipids with longer saturated acyl chains of the fatty acids and
cholesterol pack, and the more fluid liquid-disorder phase (Ld) is composed of phosphoglycerolipids with polyunsaturated fatty acids (Binder et al 2003). Lo microdomains,
called lipid rafts, function as signal transduction and protein-protein interactions to adapt
the environmental changes. Caveolae, the best characterized lipid raft, is small embolic
membrane structures (Palade 1953). Caveolae is rich in cholesterol and sphingolipids
with cholesterol-binding proteins called caveolin. Caveolins are small proteins with
several isoforms. Among these isoforms, caveolin-3 is highly expressed in myocytes, and
it is encoded by CAV3 gene (Song et al 1996, Tang et al 1996, Williams and Lisanti 2004).
The co-immunoprecipitation and immunochemistry studies for caveolin-3 indicate that
Na v 1.5 interacts and co-localizes with caveolin-3 (Yarbrough et al 2002).
1.5.3
Na v 1.5 β-Subunits
Sodium channels are composed of one pore forming α subunit and one or several
auxiliary β subunits that regulate the channel function (Meadows and Isom 2005). There
are four genes encoding human sodium channel β subunits: β1 (SCN1B) (Isom et al 1992),
β2 (SCN2B) (Isom et al 1995), β3 (SCN3B) (Morgan et al 2000), and β4 (SCN4B) (Maier
et al 2003). These four β subunits have different regulatory functions for Na v 1.5.
Coexpression of β1 and Na v 1.5 results in a current density increase (Qu et al 1995).
Coexpression of β3 and Na v 1.5 increases current density, depolarizes shifts in the
voltage-dependence of inactivation, and increases the rate of recovery from inactivation
(Fahmi et al 2001). However, anther group claimed that β3 causes hyperpolarizing shifts
20
in inactivation and slows the recovery form inactivation (Ko et al 2005). As a result, the
study about the effect of β3 on Na v 1.5 is controversial. Coexpression of β2 and β4 with
Na v 1.5 has no obvious effect on the current (Dhar Malhotra et al 2001, Yu et al 2003),
though one group found that when expressed in HEK 293 cells, a mutation (L179F) in
SCN4B increases the late sodium current, which is similar to gain-of-function mutations
in SCN5A such as ΔKPQ (Medeiros-Domingo et al 2007).
It appears that the effect of β subunits on Na v 1.5 is inconclusive and controversial
due to the possibility that the interaction between a β subunit and Na v 1.5 α subunit is
complicated and that different groups used in different cell types. β1 and β2 subunits are
associated with ankyrin, another important Na v 1.5 regulation protein discussed above.
The localization of β1-ankyrin complex is controlled by the phosphorylation of the
tyrosine on β1 carboxyl-terminus (Y181) (Malhotra et al 2002). Using a phosphorylated
antibody, tyrosine phosphorylated β1-ankyrin complex was found to colocalize with
Na v 1.5 at the intercalated disk (Malhotra et al 2004).
1.5.4
Protein Kinase A (PKA)
In cardiac cells, the β-adrenergic receptor activates a GTP-binding protein (Gs).
This GTP-binding protein then stimulates adenylyl cyclase and leads to the increase of
intracellular cAMP, which will activate PKA (Li et al 2000). It is found that PKA
activation can increase I Na by at least 30% in normal canine cardiomyocytes, but can only
increase I Na by 17% in the cells at infracted zone. Since it was hypothesized that PKA
activator could either increase the current by direct phosphorylation of the α subunit or
facilitate the channel trafficking from ER to plasma membrane (Frohnwieser et al 1997,
21
Murphy et al 1996, Zhou et al 2000), cholorquine, an inhibitor of cell trafficking, was
used to distinguish the two hypotheses. After the administration of cholorquine, the peak
I Na did not change much in infracted cells, but increased only 10% in the noninfracted
cells (Baba et al 2004). This result indicates that PKA activator increases I Na by pushing
more sodium channels to plasma membranes. Later, another group used GFP tagged
Na v 1.5 to provide direct evidence that PKA helped the trafficking of cardiac sodium
channels. They demonstrated that in the stable cell line expressing GFP- Na v 1.5, more
sodium channels localized onto the membrane, and the peak I Na increased after applying
PKA activators (Hallaq et al 2006).
1.5.5
Tyrosine Phosphorylation of Na v 1.5
Many
ion
channels
are
the
targets
for
protein
phosphorylation
and
dephosphorylation, which then alter the electrophysiological properties and activities of
the excitable cells. It was proved that voltage-gated sodium channel α subunit can be
negatively regulated by PKA phosphorylation on Ser573 located on the intracellular loop
between DI and DII (Cantrell et al 1997). Sodium channel α subunit can also be regulated
by PKC so that the peak current was reduced by a negative shift of steady-state
inactivation in the hyperpolarized direction (Qu et al 1994).
The study of tyrosine phosphorylation is more recent. One group suggested that
cardiac sodium channels can be phosphorylated by Src family tyrosine kinases. Using the
anti-phosphotyrosine antibody, the immunoprecipitated Na v 1.5 was increased in cells
expressing constitutively activated Src family tyrosine kinase Fyn. The phosphorylated
residue (Y1495) in Nav1.5 was also found (Ahern et al 2005). Another group studied the
22
effect of tyrosine kinases on INa in guinea pig ventricular myocytes. In their study, I Na
increased when being treated with epidermal growth factor (EGF). The increased I Na can
be inhibited by epidermal growth factor receptor (EGFR) kinase inhibitor tyrphostin
AG556 and enhanced by orthovanadate (a protein kinase phosphatase inhibitor) (Liu et al
2007). Jespersen T. et al found that Na v 1.5 interacts with protein tyrosine phosphatase
PTPH1 and concluded that tyrosine phosphorylation destabilizes the inactivated state of
Na v 1.5 (Jespersen et al 2006).
1.5.6
Glycerol-3-Phosphate Dehydrogenase Like Protein (GPD1-L)
In 2002, a locus on chromosome 3 was described in a large family with BrS, but no
mutation was found in SCN5A gene, which is located nearby (Weiss et al 2002). Later, a
missense mutation in GPD1-L, the glycerol-3-phosphate dehydrogenase like protein, was
found in this BrS family. As a result, it brought attentions about the interaction between
GPD1-L and SCN5A which is an important BrS gene. When co-expressed with SCN5A in
HEK293 cells, this BrS mutation in GPD1-L can reduce I Na (London et al 2007). Three
more mutations in GPD1-L were found in some patients died of sudden infant death
syndrome (SIDS). The expression of these mutant GPD1-L variants in mouse neonatal
cardiomyocytes decreases INa (Van Norstrand et al 2007).
1.5.7
Calmodulin (CaM)
Calmodulin (CaM) is a Ca2+ sensor protein and can undergo a Ca2+-induced
conformational change. This small ubiquitous intracellular Ca2+-binding protein with
only 147 amino acids participates in numerous signaling pathways and cellular processes
23
such as growth, proliferation and movement (Chin and Means 2000). During the cardiac
excitation-contraction coupling, Ca2+ handling =is crucial for many physiological events
(Saimi and Kung 2002). When interacting with other proteins, CaM usually binds to the
isoleucine-glutamine (IQ) CaM binding motif on these proteins (Herzog et al 2003). In
2002, Tan HL and co-workers found the IQ motif on the C-terminus of Na v 1.5. The
binding of CaM significantly potentiates slow inactivation (Tan et al 2002). It is also
found that Ca2+ concentrations can modulate Na+ channel through CaM (Kim et al 2004),
but recent studies suggest that this modulation is caused by the Ca2+ binding motif on the
proximal part of Na v 1.5 C-terminus (Shah et al 2006, Wingo et al 2004).
1.5.8
Ca2+/Calmodulin-Dependent Protein Kinase II (CaMKII)
CaMKII, a serine/threonine protein kinase, is activated by intracellular [Ca2+]
changes. There are four CaMKII isoforms: α, β, δ, and γ, and each isoform is encoded by
a specific gene. All the CaMKII isoforms have similar core structures including a
regulatory domain, an N-terminal catalytic domain, and a C-terminal association domain
(Braun and Schulman 1995). CaMKIIδ is predominantly expressed in the heart
(Couchonnal and Anderson 2008). Using immunoprecipitation and immunostaining, it
has been confirmed that CaMKIIδ c interacts and co-localizes with Na v 1.5 in mouse
myocytes. By overexpressing CaMKIIδ c in rabbit myocytes, Stefan Wagner et al.
successfully proved that CaMKIIδ c was able to enhance the steady-state inactivation of
I Na (Wagner et al 2006). The mice overexpressing CaMKIIδ c in the hearts developed
chronic heart failure and episodes of ventricular tachycardia. Not only does CaMKII
interact with Na v 1.5 and regulate I Na in the myocytes, it is also an important protein
24
linking Ca2+ and Na+ in the cardiac cells. As a result, CaMKII is a potential drug target
for arrhythmia treatment (Abriel et al 2000).
1.5.9
MOG1
MOG1 was first identified as a multi-copy suppressor of the yeast nuclear transport
factor GSP1 or scRan. It is a small protein encoded by RANGRF gene with molecular
weight of 29 kDa (Marfatia et al 2001, Oki and Nishimoto 1998)). Yeast two-hybrid and
in vitro binding experiments indicate that MOG1 binds to the GTP-binding nuclear
protein Ran. As a result, MOG1 is thought to regulate nuclear protein trafficking
(Marfatia et al 2001). In 2008, Ling Wu et al. found that MOG1 interacts with Na v 1.5 on
the intracellular loop between DII and DIII by a yeast two hybrid screen followed by
GST-pull down assays and co-immunoprecipitation. In HEK293 cells stably expressing
Na v 1.5, transfection of MOG1 can increase the Na v 1.5-mediated peak current without
affecting any kinetic changes of sodium channel. These results suggested that MOG1 is a
helper protein for Na v 1.5 transporting to the plasma membrane (Wu et al 2008).
1.5.10 Fibroblast Growth Factor Homologous Factor 1B (FHF1B or FGF12-1b)
Fibroblast growth factor (FGF) family is one of the largest families of polypeptide
growth factors functioning in both the adult and the embryos (Olsen et al 2003). One of
the family members, FGF12-1b (also known as FHF1B (fibroblast growth factor
homologous factor 1B)), is expressed in the heart (Abriel et al 2000). It is reported that
FGF12-1b interacted with Na v 1.5 at intracellular C-terminus motif of amino acid
1773-1832 (Goetz et al 2009, Liu et al 2003). Several LQTS and BrS mutations such as
25
E1784K, D1790G, and Y1795H and Y1795C in SCN5A are located in this FGF12-1b
binding motif. In HEK293 cells co-expressing FGF12-1b and Na v 1.5, the steady-state
inactivation curve shifted toward hyperpolarized value without any change in the peak
current density. However, FGF12-1b does not modulate or interact with D1790G mutant
Na v 1.5 (Liu et al 2003).
1.5.11 14-3-3η Protein
14-3-3 proteins are highly conserved proteins with a molecular weight of 28-33
kDa in all eukaryotic organisms (Aitken 2006). Severn family members of 14-3-3
proteins are found in mammals – β, γ, ε, σ, ζ, τ and η (Morrison 2009). The yeast
two-hybrid and co-immunoprecipitation experiments indicate that 14-3-3η interacts with
the intracellular N-terminus of Na v 1.5. The immunostaining in COS-7 cells expressing
human Na v 1.5 and 14-3-3η demonstrated that these two proteins were co-localized at the
intercalated disk of myocytes. The same result was confirmed in isolated rabbit
cardiomyocytes. Co-expression of 14-3-3 did not change the peak sodium current in
COS-7 cells, but shifted the inactivation curve towards negative potential and delayed the
recovery from inactivation. This result indicates that 14-3-3 does not regulate the
trafficking of Na v 1.5 but modified its kinetics (Allouis et al 2006).
1.5.12 Ubiquitin-Protein Ligases of the Nedd4/Nedd4-Like Family
Intracellular and extracellular protein degradation processes are different. During
extracellular degradation, the extracellular proteins are engulfed by vesicles and then
fused with primary lysosomes (Glickman and Ciechanover 2002). However, during the
26
intracellular protein degradation, it undergoes the two-steps ubiquitin-proteasome
proteolytic pathway: 1) tagging of the proteins with ubiquitin molecules, and 2)
degradation of the tagged proteins by the 26S proteasome complexes. Ubiquitin can be
recycled during this process (Glickman and Ciechanover 2002). Besides protein
degradation, ubiquitin can also regulate protein trafficking to cell membranes (Staub and
Rotin 2006).
Ubiquitin is a small protein with only 76 amino acid residues (Hershko and
Ciechanover 1998). The Nedd4/Nedd4-like family of E3 ubiquitin-protein ligase contains
a 40-residue domain that binds to various proteins with their PY motifs (Rotin and Kumar
2009). Almost all voltage-gated sodium channels have this PY motif on their intracellular
C-terminus except for Na v 1.4, Na v 1.9, and Na x (Fotia et al 2004). It is observed that I Na
was reduced by Nedd4-2 mediated ubiquitylation in Xenopus oocytes (Abriel et al 2000).
Since the total Na v 1.5 protein did not decrease in the presence of Nedd4-2, it is unlikely
that the reduction of I Na was Na v 1.5 protein degradation by Nedd4-2. Instead, Na v 1.5
proteins are trapped inside the cells by Nedd4-2 rather than pushed to the membrane
(Rougier et al 2005).
1.6 Anti-Arrhythmic Medications and Their Side Effects.
1.6.1
Quinidine
Quinidine, a class I anti-arrhythmia drug, used to be one of the most frequently
prescribed anti-arrhythmic drugs, but has side effects that may lead to arrhythmias (Grace
and Camm 1998). The pharmacokinetic study of quinidine revealed that the half life of
quinidine in humans ranges between 3 and 19 hours (Grace and Camm 1998). In cardiac
27
cells, quinidine can block fast inward sodium currents and various potassium channels
(Balser et al 1991, Wwidmann 1955). The anti-arrhythmic effect of quinidine was
thought to be caused by the inhibition of the repolarizing delayed rectifier current, which
is composed of three components including slow, rapid and ultrarapid currents (Deal et al
1996, Wang et al 1995c). Quinidine functions to inhibit the rapidly activating component
(Carmeliet 1993, Woosley et al 1993).
Initially, quinidine was used for patients with atrial fibrillation (AF), and then was
widely used as anti-arrhythmic medication (Brodsky et al 1996). However, it was later
found that the application of quinidine has a potential risk of causing syncope and sudden
death, which are due to ventricular tachycardia (VT) (Selzer and Wray 1964). The
arrhythmia induced by quinidine treatment often happened right after treatment was
started, and about 1.5%-8% of patients developed torsade de pointes (TdP) with quinidine
treatment especially in patients with AF (Roden et al 1986, Roden 1994). Patients who
developed TdP when treated with quinidine usually have abnormally longer corrected QT
interval on ECG (Roden et al 1986).
1.6.2
Mexiletine
Mexiletine is another class I antiarrhythmic drug which plays an important role in
treating cardiac arrhythmias (Sheets et al 2010). Voltage-gated sodium channels possess
different types of inactivation, including fast, slow and ultra-slow (Goldin 2008). The
slow inactivating component current I Na,P is only 1% of the peak current I Na,
T
and
considered the main target for antiarrhythmic drugs such as mexiletine and ranolazine
(Saint 2008).
28
Like other antiarrhythmic drugs, mexiletine can also promote arrhythmias
(Brugada and Wellens 1988). However, a series of test of different antiarrhythmic drugs
indicated that mexiletine might have the lowest risk of proarrhythmic action (Dhein et al
1993).
1.6.3
Ranolazine
Ranolazine is an investigative anti-arrhythmia drug in phase III trials, and believed
to block four different voltage-gate sodium channels including Na v 1.1, Na v 1.5, Na v 1.7
and Na v 1.8 (Antzelevitch et al 2011). The effect of ranolazine in treating arrhythmias is
the blockage of multiple channels such as late sodium current (I Na,L ), the late rectifying
potassium channel, and the late L-type calcium channel. It is known that the inhibition of
potassium channel will prolong the action potential duration (APD) and the inhibition of
sodium channel and calcium channel will shorten the APD. The opposite effects of
ranolazine on APD explain the modest prolongation effect on QTc in patients (Reddy et
al 2010). Similar to quinidine, the QTc increase effect of ranolazine might bring the
problem of developing drug-induced TdP. However, no such symptoms were observed
during ranolazine clinical trials (Reddy et al 2010). It is reported that only less than 2%
patients experienced the common side effects of ranolazine such as dizziness, nausea,
constipation and headache (Chaitman et al 2004, Morrow et al 2007, Stone et al 2006).
Mutagenesis study demonstrated that mutation F1760A in Na v 1.5 significantly
reduced the blockage effect of ranolazine on both peak and sustained Na+ current. It is
indicated that ranolazine interacted with the previously defined local anesthetic (LA)
receptor binding site on Na v 1.5 (Fredj et al 2006). One group reported that ranolazine
29
blocked Na v 1.5 through modulating the mechanosensitivity of this voltage-gate channel
by partitioning the lipid bilayer of the membrane (Beyder et al 2012). Na v 1.5 is
mechanosensitve in both cardiac myocytes and a heterologous system such as transfected
HEK cells, and is regulated by cytoskeleton and membrane lipid bilayer elasticity
(Lundbaek et al 2004, Undrovinas et al 1995). The interruption of lipid bilayer inhibits
the mechanosensitivity of Na v 1.5 (Beyder et al 2012).
1.6.4
Other Anti-Arrhythmic Drugs
Although newer drugs were developed and showed more clinically effective for
cardiac arrhythmia treatment, the side effect of these newer drugs brought concerns to
physicians. These newer drugs include flecainide, encainide, and moricizine, and were
found to cause ventricular premature beats after myocardial infarction and lead to
mortality in patients (Echt et al 1991). The lethal side effects brought attention to
scientists, and the study of how to reduce the side effects of the medications could be
more important than alleviating the symptoms.
1.7 Cardiac Excitation-Contraction Coupling (ECC) and Ca2+ Handling in Cardiac
Cells
1.7.1
The Importance of Ca2+ in Cardiomyocytes
Cardiac excitation-contraction coupling is the process of the contraction of the
cardiac cells to the contraction of the heart, which propels out blood to the body.
Throughout this process, Ca2+ ion plays an important role in translating the cell electrical
excitation signal to the contraction of myofilaments. Thus, the mishandling of Ca2+ may
30
cause cardiac dysfunction such as arrhythmias (Bers 2002). In the past several decades,
alterations in cytosolic Ca2+ homeostasis were considered as the main mechanism for cell
death. Ca2+ paradox is the hypercontracture and death of cardiomyocytes caused by Ca2+
restoration after a perfusion of Ca2+-free buffer (Piper 2000). Hypercontraction can cause
cardiac cell death, and the inhibition of hypercontraction using 2,3-butanedione
monoxime (BDM) during reperfusion can prevent the further damage of the cell tissues
(Garcia-Dorado et al 1992).
Ca2+ is a crucial element in all kinds of organisms, and plays an important role in
numerous biological functions. It is known that the intracellular free Ca2+ concentration
(~10-7 M) is much lower than the extracellular Ca2+. This high Ca2+ gradient provides
abundant Ca2+ available to be imported into cells, where these Ca2+ ions can function as
second messengers (Chin and Means 2000).
1.7.2
The Process of Excitation-Contraction Coupling (ECC)
In order to initiate the cardiac ECC, Na+ ions need to enter the cardiac cells through
voltage. Then, a relatively small amount of Ca2+ enters the cell through depolarizationactivated Ca2+ channels such as L-type Ca2+ channel (Figure 1). This small amount of
Ca2+ then open RYR2 on sarcoplasmic reticulum (SR) membrane and release the Ca2+
storage in SR. This process is called Ca2+ induced Ca2+ release (CICR). The rapid
intracellular Ca2+ increase allows the binding of Ca2+ to the protein troponin complex
(TN-C) which attaches to sarcomere, the myofilament repeat subunit in striated muscle
(Knollmann and Roden 2008). The binding of Ca2+ to TN-C induces a conformational
change and exposes the myosin binding sites (thick filament) on actin (thin filament).
31
L-type Ca2+ channel
Ca2+
Plasma membrane
Cytoplasm
K+
K+
ATP
Ryanodine
(Ca2+ release
channel)
Na+
Na+
Na+
Sarcoplasmic
reticulum
Ca2+
ATPase
Ca2+ uptake
pump (SERCA)
Troponin
complex
Ca2+
Na+
Na+
Na+
Myofilament
Connexin
channel
K+
Na+/Ca2+
exchanger (NCX)
Na+ channel
K+ channel
Na+
K+
K+
Figure 1. Schematic Diagram of Cardiac Excitation-Contraction Coupling (ECC) Events
in Ventricular Myocytes. In order to initiate ECCs, K+ ions enter into the cell from a
neighboring cell through connexin channels and open the voltage gated Na+ channels,
which leads to the depolarization of the membrane. The depolarization of the membrane
causes Na+ channel inactivation and opens both K+ channels and L-type Ca2+ channels.
The entry of Ca2+ ions releases sarcoplasmic reticulum (SR) Ca2+ into cytosol. This rapid
increase of cytosolic Ca2+ ions will bind to the troponin complex and activate the
myofilament contraction. During the repolarization process, the cell will remove
cytosolic Ca2+ either through Na+/Ca2+ exchanger (NCX) or Ca2+-uptake pumps on SR.
The cytosolic Na+ is excluded through the Na+/K+ pump.
32
Following the conformational change leads to movement and binding between myosin
and actin and results in the muscle contraction. This is the physiological process that how
the electrical stimulus is translated to the striated muscle contraction (Sandow 1952).
1.7.3
Cardiac Ion Channels and Cardiac Action Potential
In a cardiac cell, an action potential (AP) refers to the rapid rise and fall of a cell
electrical membrane potential. The cardiac action potential is electrical activity in the
heart and is crucial for the electrical conduction in the heart (Grant 2009). The AP
contains five phases (Figure 2), including the upstroke (0), early repolarization (1),
plateau (2), final repolarization (3) and resting phases (4), all of which are generated by
different selective ion channels on the membranes (Table III) (Grant 2009).
1.7.4 The Cellular Relaxation and the Removal of Cytosolic Ca2+
In order to exclude the increased Ca2+ in cytosol, Ca2+ ions have to be transported
out of the cell or be pumped back to SR in order to allow cell relaxation (Bers 2002). A
SR Ca2+-ATPase called SERCA on SR membrane pumps the rapidly increased Ca2+ back
into SR. Na+/Ca2+ exchanger (NCX) allows three Na+ enter the cell and exclude one Ca2+
outside the cell.
This is the main path for Ca2+ transport out of the cytosol (Pogwizd et
al 2001). There are two more pathways to exclude extra Ca2+: sarcolemmal Ca2+-ATPase
and mitochondrial Ca2+ uniporter (Bassani et al 1994).
1.7.5 Dysfunction of Na+/Ca2+ Exchanger Causes Ca2+ Overload and Arrhythmias
Na+/Ca2+ exchanger (NCX), an important protein linking Ca2+ and Na+ in the cell,
33
Phase 1
Ito
IKur
Phase 2
INa ICa,L
0 mV
IKs
INa
Phase 0
Phase 3
IKr
IK1
Phase 4
APD
Resting membrane potential -80 mV
Figure 2. A Schematic Diagram Showing that Different Ionic Currents Contribute to
Different Phases of a Cardiac Action Potential. During phase 0, the rapid inward sodium
current I Na initiates the upstroke of the action potential through Na v 1.5 channels. This
phase is followed by the inactivation of Na+ channels and the efflux of K+ ions. Next, the
action potential enters into a plateau phase (2) in which the inward movement of Na+ and
Ca2+ and the outward movement of K+ are balanced. The final repolarization phase 3 is
achieved by fast delayed rectifier, which is an outward K+ current.
34
Table III. Ionic Currents and Different Phases of AP
Currents
Channel Gene
AP Phase
Description
I Na
SCN5A
0 &2
Sodium current
I to
KCND2/3, KCNA4
KCNA7, KCNC4
1
Transient outward current
I Kur
KCN5A, KCNC1
1
Ultrarapid delayed rectifier
I Ca,L
CACNA1C
2
L-type calcium current
I NCX
NCX1.1
2
Influx of Na+ through NCX
I Kr
KCNH2
3
Fast delayed rectifier
I Ks
KCNQ1
3
Slow delayed rectifier
I K1
KCNJ2/12
4
Inward rectifier
35
is a bi-directional membrane ion transporter (Jeffs et al 2007). There are two modes in
NCX: the forward mode and reverse mode. The direction of the mode of NCX is
determined mainly by intracellular and extracellular concentrations of Na+ and
Ca2+ (Blaustein and Lederer 1999). Under normal physiological conditions, NCX
functions in the forward mode, which allows three Na+ ions enter the cell and one Ca2+
moves out of the cell (McNaughton 1991). However, during the ischemia condition, the
shortage of ATPs decreases the ability of Na+ extrusion through the energy-dependent
Na+/K+-ATPase, which further causes the intracellular Na+ overload (Murphy et al 1991).
Another study also showed that the inhibition of Na+/K+-ATPase may cause elevation of
[Na+] and SR Ca2+ overload (Sedej et al 2010). It is believed that the elevated [Na+] leads
to the reduced function of the NCX forward mode, which means that less Ca2+ ions are
excluded through NCX and therefore leads to SR Ca2+ overload (Sedej et al 2010). SR
Ca2+ overload has been experimentally proven to be the main cause for delayed
afterdepolarizations (DADs) (Lederer and Tsien 1976, Marban et al 1986). DAD is the
depolarization occurring after depolarization and related to initiation arrhythmias (Fink et
al 2011, Pogwizd and Bers 2004).
1.7.6 Intracellular Ca2+ Transient and the Development of Ca2+ Transient
Measurement
Ca2+, an important signaling ion in the cytosol, regulates the cardiac cell
contraction through binding to the Ca2+ binding protein troponin complex on
myofilaments. The affinity between Ca2+ ions and Ca2+ binding proteins depends on the
dynamic intracellular Ca2+ concentration, referred to as Ca2+ transient (Wier 1990).
36
Therefore, Ca2+ transient in striated muscle cells has provoked great interests of many
scientists.
During the development of methods for the measurement of Ca2+ transient, fura-2
was proven to be an efficient Ca2+ fluorescent indicator in individual ventricular cells
(Grynkiewicz et al 1985). Fura-2-acetoxymethyl ester (Fura-2/AM) is a membrane
permeable Ca2+ indicator. Once entering into the cells, the cellular esterases will remove
the acetoxymethyl groups on Fura-2/AM, and then Fura-2, the real working Ca2+
fluorescent indicator, is generated. The ratio of fluorescence at 340 nm and 380 nm
(340/380) is proportional to the intracellular Ca2+ concentration (Grynkiewicz et al 1985).
37
CHAPTER II.
CARDIAC EXPRESSION OF SCN5A SNP R1193Q IN TRANSGENIC MICE
PROLONGS QT INTERVAL
2.1 Abstract
Type 3 LQTS (LQT3) is associated with mutations in the cardiac sodium channel
gene SCN5A. Some LQT3 mutations generate persistent, late sodium currents (I Na,L ).
A single nucleotide polymorphism (SNP), i.e., R1193Q, was previously reported in a
patient with drug-induced LQTS and further shown to be present in 0.2% of the general
Caucasian population and nearly 12% of the Chinese population. SNP R1193Q mutation
destabilizes inactivation gating of the cardiac sodium channel and generates I Na,L that is
expected to prolong the cardiac action potential duration (APD) and the QT interval on
electrocardiogram (ECG). In order to study this mutation, our lab created a transgenic
mouse line carrying human R1193Q mutant SCN5A. Using telemetry ECG analysis, I
found that cardiac-specific expression of R1193Q results in prolongation of QT interval
38
in transgenic mice (106.3 ± 9.47 ms, n = 7) compared to control groups of wild type mice
(81.6 ± 14.12 ms, n = 5, p < 0.001) and transgenic wild type mice (78.0 ± 7.26 ms, n = 6,
p < 0.001), demonstrating that R1193Q confers risk of prolonged QT interval, which
potentially could be pro-arrhythmic. Moreover, in the presence of quinidine, R1193Q
transgenic mice showed a higher incidence or premature atrial contractions (PACs),
premature ventricular contractions (PVCs) and sinus atrial exit block (SAEB) than
littermate wild type control mice and transgenic wild type mice.
In this study, I show that cardiac-specific expression of human R1193Q mutant
SCN5A results in prolongation of QT interval in transgenic mice, which unequivocally
demonstrates that R1193Q confers risk of prolonged QT interval, which is pro-arrhythmic.
Moreover, in the presence of quinidine, R1193Q transgenic mice (TG-RQ) showed a
higher incidence of premature ventricle contractions (PVCs) or premature atrial
contractions (PACs) and AV block than littermate wild type control mice. These results
suggest that R1193Q may increase the risk for drug-induced arrhythmias. Considering the
high frequency of R1193Q in the general population, it may become necessary to perform
population-wide screening for the variant to identify high risk individuals.
2.2 Introduction
Long QT syndrome (LQTS) is an a cardiac disorder characterized by prolonged QT
interval on surface electrocardiogram (ECG), syncope and sudden death as a result of
life-threatening ventricular tachycardia, specifically torsade de pointes (Wang et al 2004).
Among them, type-3 long QT syndrome is caused by mutations in cardiac sodium
channel gene SCN5A encoding type 5 voltage gated sodium channel α subunit, Na v 1.5
39
(Wang et al 1995a). Nav1.5 is a large integral membrane protein which mediates the rapid
upstroke of action potential in cardiac cells (George et al 1995).
Acquired LQTS is generally drug-induced and frequently caused by therapeutic
drug effects on ion channels involved in genetic forms of LQTS, such as blockage of the
cardiac potassium channels (Heist and Ruskin 2010). The missense mutation R1193Q
was first identified in one of seven acquired LQTS patients. The electrophysiological
study demonstrated that this mutation destabilizes inactivation gating and generates a
persistent, non-inactivating sodium current in both transfected human kidney (HEK293)
cells and Xenopus oocytes (Wang et al 2004). This SNP R1193Q presents in 0.2% in
Caucasian general population and 12% in Chinese population (Hwang et al 2005, Wang
et al 2004). Later, one group reported that SCN5A R1193Q polymorphism in a Chinese
family was associated with progressive cardiac conduction defects and LQTS (Sun et al
2008).
Because some carriers with R1193Q have prolonged QTc, that my derivate in
acquired LQTS and some have apparently normal QTc on surface ECG, it is necessary to
demonstrate whether R1193Q confers a risk of QTc prolongation under action of drugs
that can have the potential of prolonging the QT interval. The purpose of this study is to
evaluate the ECG characteristic in transgenic mice carrying the high prevalence SNP
R1193Q mutation, and the drug effects of quinidine, a class I anti-arrhythmic agent which
also has class III properties, on this LQTS mouse model.
2.3 Materials and Methods
2.3.1 Preparation of Experimental Animals
40
This study was conducted in accordance with the guideline approved by the
Cleveland Clinic Foundation Institutional Animal Care and Use Committee and was
conformed to NIH guidelines.
The strategy for generating transgenic mice carrying SNP R1193Q in SCN5A
(TG-RQ) is similar to that used for developing transgenic mice which overexpress wild
type SCN5A (TG-WT) (Tian et al 2004)or mutant SCN5A with the N1325S mutation
(TG-NS) (Tian et al 2004, Tian and Wang 2006, Zhang et al 2011). In brief, a cDNA
fragment of 6.2 kb human SCN5A (hSCN5A) with SNP R1193Q was subcloned behind
the cardiac specific mouse myosin heavy chain-α (α-mMHC) promoter (5.4kb), and
before a 0.6 kb human growth hormone polyadenylation signal sequence (hGH plA),
resulting in a mMHC-hSCN5A construct. This construct was digested with Not I. A 12.4
kb Not I fragment was purified by agarose gel electrophoresis and injected into fertilized
eggs derived from the CBA/B6 mouse strain to yield founder mice. Founder mice were
then bred to wild type mice to generate heterozygous TG-RQ mice. The experimental
control mice in this study are non-transgenic wild type mice (denoted as NTG) or
TG-WT as described previously (Tian et al 2004, Tian and Wang 2006, Tian et al 2007,
Zhang et al 2011).
Genotyping of positive TG-RQ mice was identical to that used for identifying
TG-WT or TG-NS mice as previously described (Tian et al 2004, Tian and Wang 2006,
Zhang et al 2007, Zhang et al 2011). In brief, mouse genomic DNA was isolated from
clipped tail tissue samples when the pups were 10 days or younger. The tail tissue
samples were collected in a 1.5 ml tube by adding 600 µl freshly prepared 50 mM NaOH.
In order to fully digest the tissue samples, the mixture was heated on a heat block at 95oC
41
for one hour. After digestion, the mixture was placed on ice for 5 minutes and 50 µl of
Tris-HCl buffer (pH = 8.0) was added. The mixture contained mouse genomic DNA and
was ready to be used as DNA template for PCR-based genotyping. The primers for
genotyping of TG-RQ mice were identical to those used for genotyping TG-WT and
TG-NS mice and included 5’-TGT CCG GCG CTG TCC CTC TG-3’ (P1, forward) and
5’-CTC ATG CCC TCA AAT CGT GAC AGA-3’ (P2, reverse) (Tian et al 2004).
2.3.2 Western Blot Analysis
Western blot analysis was performed to estimate the expression level of Nav1.5 in
cardiac tissue samples from transgenic mice. Total protein extracts were extracted from
the homogenized hearts using lysis buffer containing 20 mM Tris-HCl (pH = 8.0), 100
mM NaCl, 1 mM EDTA and 0.5% NP-40. The concentration of protein extracts was
measured using the Bio-Rad protein assay kit (Bio-Rad, Hercules, CA) and all samples
were diluted to the same concentration of 10 µg/µl. The protein extracts (100 µg) were
separated by an 8% SDS-PAGE prepared with National Diagnostics ProtoGel solution
(National Diagnostics, Atlanta, GA). After the separation, proteins were transferred to
polyvinylidene difluoride membranes (Immun-blot PVDF membrane) (Bio-Rad,
Hercules, CA) at 4 oC and 40V overnight. The membrane was blocked in 5% non-fat
milk in PBST (phosphate-buffered saline, 0.05% (v/v) tween 20, pH=7.4) and then
probed with an anti-rabbit Nav1.5 antibody (1:1000 dilution), followed by incubation
with a horseradish peroxidase-conjugated secondary antibody (Santa Cruz, Santa Cruz,
CA). The protein signal was visualized using the ECL Western blotting detection kit and
Hyperfilm ECL (GE Healthcare, Pittsburgh, PA). The membrane was striped and blocked
42
with 5% non-fat milk in PBST, and was probed again using an anti-GAPDH antibody
(Millipore, Billerica, MA), which serves as a loading control.
2.3.3 Semi-Quantitative Real-Time PCR (RT-PCR) Analysis
Total RNA samples were isolated from heart tissue samples from both NTG and
TG-RQ mice in order to determine the expression level of SCN5A mRNA. Excised
mouse hearts were placed into liquid nitrogen right after being washed with PBST. Total
RNA samples were isolated using TRIzol reagent (Life Technologies, Grand Island, NY).
For each heart sample, 1 ml of TRIzol reagent was applied, and the heart tissue was
homogenized manually. After isolation, the RNA was immediately transcribed into cDNA
using Stratagene’s RT-PCR kit (Agilent Technologies, Santa Clara, CA). Double-stranded
cDNA was synthesized from 10 µg of total RNA. The semi-quantitative RT-PCR was
used to quantify the RNA level. PCR primers were designed to amplify a region from
exon 5 to 7 of the human SCN5A gene. The sequence of the forward primer is 5’- TGG
AAC TGG CTG GAC TTT AGT GTG -3’, and the reverse primer is 5’-TGG TGC CGT
TGA GCG CTG TGA AG -3’. RT-PCR analysis with 18S rRNA primers was used as
internal control. The number of cycles for semi-quantitative RT-PCR was optimized for
each sample to obtain data in the linear range for amplification. The PCR profile was 25
cycles for SCN5A and 18 cycles for 18S rRNA. The PCR products were separated by a
1.5% agarose gel and analyzed.
2.3.4 Telemetry ECG Recordings
Telemetry Electrocardiogram (ECG) Data from TG-RQ and control mice were
43
acquired using Data Sciences telemetry system (TA10ETA-F20 transmitter, Data
Sciences International) with a receiver placed under each mouse cage. Mice (25-32g,
6-10 months) were anesthetized with 20 mg/ml of Avertin (0.025 ml/g mice)
(Sigma-Aldrich, St. Louis, MO). The abdomen hair was removed and the telemetry
transmitter was then implanted into the mouse peritoneal cavity. The negative and the
positive leads on the telemetry transmitter were surgically placed on the right shoulder
and chest, respectively. The mice were observed for the first 48 hours after surgery, and
ECG data were collected at least 48 hours after surgery using Data Sciences DATA
Acquisition ECG software (Data Science International, St. Paul, MN). The QT interval
was corrected for the RR interval using Bazett’s formula by QTc = QTo/(RR)1/2 (Bazett
1920, Tian et al 2004, Zhang et al 2007) or the formula of QTc = QTo/(RR/100)1/2 as
described previously (Mitchell et al 1998).
2.3.5 Echocardiographic Assessment of Cardiac Structure and Function in
Transgenic Mice
The echocardiographic data were collected from TG-RQ and control mice (8-10
month) as described previously (Zhang et al 2006). Each mouse had at least one
echocardiography with the same handler. Echocardiography was performed using a highframe rate (> 200 fps) echocardiography machine (GE Medicare, Milwaukee, WI) with
14-MHz transducer coupled with an Applio ultrasopund machine (Toshiba, Japan).
2.3.6 Data Analysis
All the results were expressed as mean ± S.E.M., and the differences between
44
means were calculated by Student’s paired t-test. A P value of 0.05 or less is considered
to be statistically significant.
2.4 Results
2.4.1 Generation and Identification of Transgenic Mice with Cardiac Specific
Overexpression of Mutant Human SCN5A Gene with SNP R1193Q
To determine whether SNP R1193Q in cardiac sodium channel gene SCN5A
increases risk of LQTS, we generated TG-RQ transgenic mice with cardiac selective
expression of mutant SCN5A with SNP R1193Q under the control of the mouse α-myosin
heavy chain promoter, a cardiac-specific promoter (Figure 3A). Semi-quantitative realtime PCR analysis showed that TG-RQ mice expressed 5-fold more SCN5A mRNA in the
heart tissue than control NTG mice (Figure 3B). Similarly, western blot analysis showed
that TG-RQ mice expressed more than 3-fold more Na v 1.5 protein in the heart than
control NTG mice (Figure 3C). These results suggest that TG-RQ mice successfully
overexpress mutant SCN5A with SNP R1193Q in the hearts.
45
Figure 3. Generation and Genotyping of TG-RQ Mice. (A) Transgenic construct for the
engineering of transgene, RQ (SCN5A) (human SCN5A with acquired LQTS-associated SNP
R1193Q), into the mouse genome. (B) Semi real-time PCR analysis was performed using heart
tissue samples from non-transgenic mice and TG-RQ mice to estimate the SCN5A mRNA
expression level in mouse hearts. SCN5A mRNA expression level in TG-RQ mice is 5-fold more
than NTG. (C) Western blot analysis was performed using mouse hearts to estimate the protein
expression level of SCN5A. SCN5A protein expression level is 3-fold more than NTG.
46
2.4.2 Development of Long-QT Syndrome in TG-RQ Mice
I recorded ECGs from conscious and unrestrained mice using the telemetry
monitoring system. Representative ECG traces for NTG, TG-WT and TG-RQ mice are
shown in Figure 4A. The major ECG parameters were measured and compared among
different groups of mice. Average QTc in TG-RQ mice was 106.3 ± 9.47 ms (n=7), which
was significantly longer than that in either NTG mice (81.6 ± 14.12 ms, n=5) (P < 0.001)
or TG-WT mice (78.0 ± 7.26 ms, n=6) (P < 0.001) (Figure 4B). These results suggest that
SCN5A SNP R1193Q increases QTc in mice. Similar significant differences were
obtained when QTc was calculated using either the Bazett’s formula by QTc = QT/(RR)1/2
(Bazett 1920, Tian et al 2004, Zhang et al 2007) or the formula of QTc = QT/(RR/100)1/2
as described previously (Mitchell et al 1998).
The PR interval in TG-RQ mice (16.4 ± 1.55 ms) was similar to that in TG-WT
mice (18.6 ± 2.37 ms) (P > 0.05), but significantly shorter than that in NTG mice (31.5 ±
2.70 ms, P < 0.001) (Figure 4B). The QRS complex in all three groups of mouse lines
had no significant differences (Figure 4B).
47
Figure 4. ECG Recordings of Non-transgenic (NTG), Transgenic Wild-type (TG-WT)
and Transgenic R1193Q (TG-RQ) Mice. (A) Representative sample ECG traces from
non-transgenic (NTG), TG-WT and TG-RQ mice. (B) Summary of ECG data analysis.
QT intervals were corrected for heart rate as indicated as QTc. Statistical significance
from control is denoted with an asterisk (*), whereas non-significance is indicated as NS.
48
2.4.3 Quinidine Prolonged QTc, and Increased the Frequencies of PACs, PVCs and
Sinus Atrial Exit Block in TG-RQ Mice
SCN5A SNP R1193Q was identified in an acquired LQTS patient upon
administration of quinidine(Wang et al 2004). Therefore, I tested whether quinidine could
further increase QTc in TG-RQ mice. TG-RQ and control NTG mice were treated with 25
mg/kg body weight of quinidine (i.p.), and ECGs were recorded using the telemetry ECG
monitoring system. As shown in Figure 5, quinidine treatment significantly increased
QTc in both TG-RQ and NTG mice 5 minutes after quinidine treatment, although
quinidine diminished the QTc differences between these two groups. After quinidine
treatment, QTc increased from 106.3 ± 9.47 ms to 128.1 ± 5.7 ms (n = 7, P < 0.05) in
TG-RQ mice, and from 81.6 ± 14.12 ms to115.7 ± 6.2 ms (n= 5, P < 0.01) in NTG mice.
TG-RQ mice remained to have significantly longer QTc 15 minutes and 25 minutes after
treatment compared to control mice.
49
Figure 5. Quinidine Treatment Prolonged QTc in Both NTG (n=5) and TG-RQ Mice
(n=7). (A) Representative ECG traces before and after quinidine treatment in both NTG
and TG-RQ mice. (B) Comparison of QTc before and after quinidine treatment.
Quinidine treatment significantly prolonged QTc in both NTG and TG-RQ mice. *, P <
0.05; **, P < 0.01; NS, not significant.
50
Representative ECG traces in TG-RQ mice after quinidine treatments are shown in
Figure 6. The analysis of ECGs showed frequent premature atrial contractions (PACs),
premature ventricular contractions (PVCs) and sinus atrial exit block (SAEB, or sinus
node pause). The frequencies of PACs, PVCs and SAEB one hour before and after
quinidine treatments were summarized in Table IV. Before quinidine treatments, no NTG
mice (n = 5) showed any pro-arrhythmic ECG patterns, whereas three of seven TG-RQ
mice showed PACs, PVCs, and SAEBs. However, after the application of quinidine, one
NTG mouse showed a low frequency of PACs/PVCs, whereas five of seven TG-RQ mice
developed PACs and PVCs. The frequencies of PACs and PVCs were much higher after
quinidine treatments than before the treatments in TG-RQ mice. Moreover, three of seven
TG-RQ mice developed SAEBs more than 25 times in one hour.
51
Figure 6. Quinidine Treatment Induced Increased PVCs, PACs, and SAEB in TG-RQ
Mice.
52
Table IV. Effects of Quinidine Treatment on PACs, PVCs and SAEB in TG-RQ Mice
and Control NTG Mice
Before Quinidine
After Quinidine
Mouse ID#
Number of PACs
and PVCs
Number of PACs
and PVCs
Number of SAEB
NTG 1
0
2
0
NTG 2
0
0
0
NTG 3
0
0
0
NTG 4
0
0
0
NTG 5
0
0
0
TG-RQ 1
0
4
26
TG-RQ 2
2
6
0
TG-RQ 3
2
9
0
TG-RQ 4
2
16
0
TG-RQ 5
0
5
2
TG-RQ 6
0
0
> 30
TG-RQ 7
0
0
> 30
Notes: The number of PACs, PVCs and SAEB in a period of one hour before and after
quinidine treatment was counted manually and summarized.
53
2.4.4 Echocardiographic Assessment of Cardiac Structure and Function in the Mice
Because TG-NS mice developed dilated cardiomyopathy and heart failure at
advanced ages as in human carriers with SCN5A mutation N1325S (Zhang et al 2011),
echocardiography for TG-RQ mice was also performed, too. Basic echocardiographic
parameters were compared between TG-RQ mice and control NTG mice (Table V).
There were no significant structural differences between TG-RQ mice and NTG mice
with regard to left ventricular fractional shortening (LVFS), systolic left ventricular end
dimension (LVEDS), systolic intraventricular septum thickness (IVSS), left ventricular
systolic posterior wall thickness (LVPWS), and other major echocardiographic
parameters (Table V). These results suggest that SCN5A SNP R1193Q does not cause
major structural changes in the heart.
54
Table V. Echocardiographic Assessment of Cardiac Function in TG-RQ Mice
Parameter
NTG (n=7)
TG-RQ (n=7)
P Value
HR (bpm)
638.3 ± 0.7
672.4 ± 0.5
0.59
IVSD (mm)
1.1 ± 0.08
0.9 ± 0.03
0.14
IVSS (mm)
1.8 ± 0.08
1.6 ± 0.04
0.09
LVIDD (mm)
3.3 ± 0.1
3.1 ± 0.1
0.41
LVIDS (mm)
1.5 ± 0.1
1.5 ± 0.1
0.8
LVPWD (mm)
1.0 ± 0.09
0.9 ± 0.06
0.65
LVPWS (mm)
1.5 ± 0.08
1.5 ± 0.05
0.73
LVFS(%)
57.0 ± 2.4
51.0 ± 3.3
0.18
Notes: HR - heart rate; IVSD - left ventricular septum wall thickness in diastole;
IVSS-left ventricular septal wall thickness in systole; LVEDD - left ventricular cavity
size in diastole; LVEDS - left ventricular cavity size in systole; LVPWD - left ventricular
posterior wall thickness in diastole; LVPWS - left ventricular posterior wall thickness in
systole; LVFS - left ventricular fractional shortening, percent change in left ventricular
cavity dimensions with systolic contraction. The mice used for echocardiography were at
the age of 8 - 10 months.
55
2.5 Discussion
SNP R1193Q of SCN5A is a relatively high frequent variant in the general
population, accounting for 0.2% the Caucasian population and 12% of the Chinese
population (Hwang et al 2005, Wang et al 2004). SNP R1193Q was identified in one
patient with drug-induced LQTS and follow-up electrophysiological studies demonstrated
that it had similar functional effect on cardiac sodium channels as other types of
LQTS-causing mutations such as N1325S. R1644H, and DelKPQ (Huang et al 2006,
Wang et al 2004). Specifically, R1193Q can generate a late non-inactivation sodium
current through an increased rate of dispersed reopening, which is predicted to prolong
the cardiac action potential duration and QTc on ECGs (Huang et al 2006, Wang et al
2004). However, genetic evidence for the association between SNP R1193Q and
prolongation of QTc was not particularly strong. In this study, our lab established TG-RQ
transgenic mice with cardiac specific expression of mutant human mutant SCN5A gene
with the R1193Q SNP in mice. Telemetry ECG recordings showed that SNP R1193Q
significantly increase QTc in TG-RQ mice compared to comparable control TG-WT or
NTG mice (Figure 4). These data provide strong in vivo genetic data to support the
hypothesis that SNP R1193Q of SCN5A can prolong QTc on surface ECGs.
One carrier with SNP R1193Q developed acquired LQTS after administration of
quinidine (Wang et al 2004). In this study, we found that quinidine increased QTc from
TG-RQ mice (Figure 5). These data support that data from the human patient that
quinidine can further prolong QTc from R1193Q carriers. Because quinidine increased
QTc from both TG-RQ mice and control NTG mice (Figure 5), the relationship between
genotype R1193Q and quinidine is additive with regard to the QTc phenotype, meaning
56
no epigenetic interaction between the two. We previously reported that although
quinidine inhibited the peak sodium current form both wild type and R1193Q mutant
sodium channels, but did not block the persistent late sodium current, which is more
relevant to prolongation of QTc (Wang et al 2004). Therefore, quinidine may prolong
QTc in TG-RQ mice or NTG mice by blocking outward potassium currents. The data
from TG-RQ mice reinforce the notion that acquired LQTS represents a latent form of
inherited LQTS because they share the similar genetic basis, but may manifest to the full
LQTS upon additive effects of quinidine and other medications as well as other
environmental factors (Wang et al 2004). A precaution should be taken when prescribing
quinidine or other QT-prolonging drugs to patients carrying SNP R1193Q.
In addition to in an acquired LQTS patient, SNP R1193Q was also identified in
several patients with typical LQTS, Brugada syndrome, cardiac conduction disease, and
sudden unexpected nocturnal death syndrome, and in a South Korean patient with LQTS,
dilated cardiomyopathy, and sinus node pause (Huang et al 2006, Hwang et al 2005,
Kwon et al 2012, Matsusue et al 2012, Qiu et al 2009, Skinner et al 2005, Sun et al 2008).
Interestingly, I observed frequent SAEB or sinus node pause after administration of
quinidine (Figure 6), a clinical feature observed in a Korean patient. No features related
to dilated cardiomyopathy and Brugada syndrome were observed. No polymorphic
ventricular tachyarrhythmia and sudden death were detected in TG-RQ mice, although an
increased frequency of PACs and PVCs were found after administration of quinidine
(Table IV). It remains to be determined whether human carriers with SNP R1193Q have
increased PACs or PVCs.
There are a few limitations with the present study. The first limitation is related to
57
the transgenic overexpression technology. Because LQTS-associated SCN5A mutations
are gain of functions mutations including R1193Q that generate a persistent late sodium
current, it is justified to overexpress R1193Q to investigate its effect on cardiac
physiology. However, one disadvantage is that the observed phenotypes may be related to
overexpression of sodium channels. For example, the shortened PR interval, development
of PACs and PVCs, and depressed T waves may be due to overexpression of sodium
channels because these phenotypes were observed in TG-WT mice, too (Zhang et al
2007). Although utilization of TG-WT mice as control may eliminate these non-mutation
specific effect, these factors need to be considered during interpretation of the data.
Second, transgenic overexpression studies may exaggerate the mutant phenotype due to a
higher number of the mutant SCN5A than the number of endogenous wild type SCN5A.
Future studies with a knock-in model of R1193Q may be needed to mimic the human
situation where there is only one copy of mutant SCN5A and one copy of the wild type
gene.
In conclusion, cardiac specific overexpression of mutant SCN5A with SNP1193Q
led to increased QTc in transgenic mice (TG-RQ) compared to mice with cardiac specific
overexpression of wild type SCN5A or NTG. Quinidine treatment can further increase
QTc in TG-RQ, although it also increased QTc in control NTG control. Our data provide
strong genetic evidence that R1193 can increase risk of LQTS and quinidine can further
exaggerate the QT phenotype. Considering that 60,000 Americans and 156 million
Chinese people are potential carriers of SNP R1193Q, this functional variant may be a
significant risk factor for LQTS in the general population, in particular, under conditions
of other QT-prolonging medications.
58
CHAPTER III
LQTS MUTATION N1325S IN SCN5A CAUSES DISRUPTION OF CYTOSOLIC
CALCIUM HOMEOSTASIS IN MOUSE VENTRICULAR MYOCYTES
3.1 Abstract
The gain-of-function mutation N1325S in cardiac sodium channel gene SCN5A
produces the late sodium current (I Na,L ) and is associated with type 3 long QT syndrome
(LQT3), ventricular tachyarrhythmia (VT) and dilated cardiomyopathy (DCM) in both
human and mice. An increased rate of apoptosis was also found in the hearts of transgenic
mice expressing N1325S mutant human SCN5A (TG-NS). We hypothesized that INa,L
their recoveries from caffeine applications, and Na+/Ca2+ exchanger (NCX) dysfunction
in isolated TG-NS myocytes. Then, it was concluded that the loss of NCX activity
confines cytosolic Ca2+ removal to SR Ca2+ reuptake and, thereby, prolongs Ca2+ recovery.
To our knowledge, this is the first time to link the sodium channel mutation to abnormal
Ca2+ handling. But there've been brought a lot of doubts about the linkage between I Na,L
59
and the abnormal Ca2+ transient by some reviewers. In order to further support the
linkage between gain-of-function sodium channel mutations with abnormal Ca2+ handling,
I have applied a late sodium current blocker, ranolazine, to isolated TG-NS myocytes in
this study. The blockage of I Na,L by ranolazine markedly reduced irregularity of the Ca2+
transients and the recovery time from caffeine effect (indicated as T90r) in TG-NS
myocytes. The rescue effect of ranolazine to abnormal Ca2+ handling provides additional
evidence that a sodium channel mutation is the cause for abnormal Ca2+ handling in
cardiomyocytes, which provides important insights into the pathogenic mechanisms of
LQTS, VT and DCM.
3.2 Introduction
In cardiac cells, the transient opening of Na+ channels gives rise to the inward
current which is responsible for the upstroke of the cardiac action potential (AP) and
initiation of excitation-contraction coupling (Li et al 1997, Liu et al 1992). After opening,
most Na+ channels quickly enter an inactivated state. However, a small fraction of Na+
channel either fails to inactivate or to reopen during the plateau of the AP when the
channels are normally closed. As a result, the outside Na+ ions continue to flow into the
cell, and the influx of Na+ produces persistent or late Na+ currents (I Na,L ). Then, I Na,L
slows the repolarization of the AP and therefore increases the action potential duration
(APD), proliferating heterogeneity of repolarization, and promoting the formation of
early after-depolarizations (Kiyosue and Arita 1989).
It is known that under an ischemia and/or reperfusion condition, where Na+
channel activation is impaired, I Na,L contributes to a rise of myocardial Na+ content in the
60
cells (Barrington et al 1997, Williams and Lisanti 2004). The increase of intracellular
[Na+] can lead to intracellular Ca2+ overload via the dysfunction of Na+/Ca2+ exchanger
(NCX), and these observations are consistent with the findings that an enhanced I Na,L can
lead to a Na+-dependent Ca2+ overload (Fraser et al 2006, Pieske et al 2002, Song et al
2006, Wang et al 2008).
N1325S mutation in the cardiac sodium channel gene SCN5A is associated with
type 3 long QT syndrome (LQT3), which was originally identified in a 96-member
family in 1995(Wang et al 1995a). In order to study this mutation, our lab has created a
transgenic mouse model in two different lines (TG-NSL3 and TG-NSL12) that
overexpress this NS mutation in the heart with different copy numbers. At the same time,
a control model in two different lines (TG-WTL5 and TG-WTL10) that carry a similar
copy number of wild type SCN5A in the heart was also generated. The resulting TG-NS
phenotype of prolonged APDs in isolated ventricular cells, increased QTc, a high
incidence of both ventricular tachyarrhythmias (VT) and sudden death has been reported
previously (Tian et al 2004). Then follow-up studies in these TG-NS mice revealed
co-development of dilated cardiomyopathy (DCM) and increased apoptosis and fibrosis
(Zhang et al 2011). After these findings, in order to determine the mechanistic basis for
this association, the Ca2+ handling was tested in isolated ventricular myocytes. The
existence of cytosolic Ca2+ irregularities from TG-NS ventricular myocytes but not from
TG-WT hearts is consistent with previous results that the heterogeneity of APD
waveforms was effectively reduced in the presence of the L-type Ca2+ channel blocker,
verapamil (Yong et al 2007). The major findings in isolated TG-NS ventricular myocytes
include increased intracellular [Na+], Na+/Ca2+ exchanger dysfunction, and abnormal
61
Ca2+ handling. These findings are the first to demonstrate a direct link between SCN5A
mutations and abnormal Ca2+ handling in cardiomyocytes. This potential link offers
possible mechanisms into the high rate of apoptosis and fibrosis in TG-NS mice heart
tissue and the existence of extracardiac diseases such as DCM. However, the
up-regulation of some important Ca2+ handling proteins such as NCX, RyR2 and L-type
Ca2+ channel in adult TG-NS mice hearts indicates the remodeling of the cardiomyocytes
in the presence of N1325S mutation. In order to strengthen the direct linkage between
this gain-of-function SCN5A mutation and abnormal Ca2+ handling, a late sodium current
blocker, ranolazine (Antzelevitch et al 2011, Fredj et al 2006, Zaza et al 2008), was
applied to the isolated myocytes. The rescue effect of ranolazine to the abnormal Ca2+
transient signals in TG-NS myocytes built a stronger bond between N1325S mutation and
cytosolic Ca2+ irregularity. The possibility that the Ca2+ mishandling is caused by the
remodeling of adult TG-NS myocytes is also diminished due to the ranolazine rescue
effect.
3.3 Methods and Materials
All experiments were conducted in accordance with the guidelines of the Cleveland
Clinic Foundation Institutional Review Board on Animal Subjects and conformed to NIH
guidelines. The generation and genotyping of TG-NSL3, TG-NSL12, TG-WTL5 and
TG-WTL10 transgenic mice have been described previously (Tian et al 2004, Zhang et al
2007).
3.3.1 Isolation of Mouse Ventricular Myocytes
62
The isolation procedure of ventricular myocytes is similar to what has been
previously reported (Tian et al 2004, Yong et al 2007). The mouse was heparinized with
150 U (0.15ml of 1000 U/ml) of heparin via intraperitoneal (IP) injection 20 minutes
before neck dislocation sacrifice. Then the mouse chest was quickly opened, and the heart
was removed and washed with ice cold surgery buffer containing (in mM): 200, NaCl; 10,
KCl; 5, MgCl 2 ; 2, KH 2 PO 4 ; 22, glucose; 50, HEPE; 4, CaCl 2 ; 12.5, pyruvic acid; pH
7.25 with NaOH. Next, under the microscope, the aorta of the heart was mounted on a
syringe needle for Langendorff apparatus with a constant perfusion buffer flow rate of
1.5ml/minute, and the aorta was secured with silk suture. The procedure should be
finished within 5 minutes after the mouse was sacrificed. Before surgery, the Langendorff
perfusion system was filled with perfusion solutions and cleared of air bubbles in order to
prevent bubbles from entering the aorta during perfusion. The perfusion solutions were
saturated with 95% O 2 :5% CO 2 and warmed to 37oC using a heated water bath. The
cannulated heart was first washed with 100 ml first step Ca2+ containing perfusion buffer
containing (in mM) 118, NaCl; 4.8, KCl; 2, CaCl 2 ; 2.5, MgCl 2 ; 1.2 KH 2 PO 4 ; 11, glucose;
13.8, NaHCO 3 ; 4.9, pyruvic acid; PH 7.2. During this stage, a healthy and successful
perfused heart should exhibit strong contraction and the blood in the heart chamber
would be flooded out. After 3-minutes of washing with Ca2+ containing perfusion buffer,
the heart was transferred to another perfusion system containing 60 ml second step
Ca2+-free perfusion buffer. Then 45 mg of type II collagenase (Worthington, Lakewood,
NJ) was added to second step Ca2+-free perfusion buffer once the heart stopped beating.
Three minutes after adding collagenase, 1 mg of protease (Type XXIV, Sigma-Aldrich, St.
Louis, MO) was added. Two sets of 0.1 mM CaCl 2 was added 5 minutes (5 µl) and 10
63
minutes (10 µl) after applying collagenase. About 25 minutes after second step perfusion,
the ventricle should appear yellowish and soft, and the digestion perfusion completed.
The heart was placed in warm (~35oC) KB media buffer to reduce “Ca2+ paradox”
(Isenberg and Klockner 1982), and the atrial appendages were carefully removed. Gentle
tituration was performed in order to loosen the cells from the heart tissue. After tituration,
a series of incremental increases in Ca2+ was made to a final Ca2+ concentration of 1.8
mM. The cells are rested at room temperature for one hour before recordings.
The cells were split into two sets in order to test the rescue effects of ranolazine.
The ranolazine was dissolved in PBST:DMSO (95:5) to 10 mM and saved as stock. One
set of cells were incubated in 10 µM of ranolazine when “resting”, the other set of cell
were incubated with same amount of blank PBST:DMSO (95:5) solvent that was used to
dissolve ranolazine.
3.3.2 Measurements of Ca2+ Transient Signals
The Ca2+ transient signals were recorded according to the protocols reported
previously (Lalli et al 2001, Ritter et al 2000, Rota et al 2005). Myocytes were loaded
with 1 µM fura-2 acetoxymethyl ester (Fura-2-AM, Teflabs, Austin, TX) for 10 minutes
at room temperature in the dark. Then the cells were centrifuged, and the buffer
containing fura-2 was removed. Next, the cells were re-suspended in 3-4 ml of HEPE
buffer and put in the dark.
Before measurement, cells were transferred to the Bioptechs chamber (Bioptechs,
Butler, PA) for which the temperature was set at 28oC. The initial field stimulation was
set at 0.5 Hz and 5 ms duration. Ca2+ transients were measured using the same excitation
64
(340/380 nm) and emission (510 nm) spectra from the spectofluorometer. As an index of
Ca2+ changes, the ratio of the emitted fluorescence at the wavelength of 340 nm over 380
nm (F/F o ) was obtained. The steady-state Ca2+ transient signals were recorded at a pacing
frequency at 0.5 Hz in the absence of caffeine. The relative sarcoplasmic reticulum (SR)
Ca2+ contents were estimated by the amplitudes of Ca2+ transient after the application of
10 mM caffeine. The stimulation was stopped before adding caffeine. The resulting
caffeine-induced Ca2+ transient (the amplitude from baseline to peak of the signal) was
considered as a qualitative index of the SR Ca2+ content. The time for a caffeine-induced
Ca2+ transient to recover by 90% from peak to baseline (T90r) was considered as the
effectiveness of the SR and NCX in the myocytes to remove cytosolic Ca2+ back to SR or
out of the cytoplasm.
3.4 Results
Irregular Ca2+ Transients in TG-NS Cardiomyocytes Can be Rescued by Late
Sodium Current Blocker Ranolazine
According to the intracellular [Na+] measurement done by Dr. Yong in our lab, it
was confirmed that the intracellular [Na+] in TG-NS myocytes was higher than in
TG-WT myocytes. An increase in [Na+] can potentially lead to Ca2+ overload in the cells,
perhaps through disrupting the cell’s ability to internally regulate Na+ and Ca2+ ionic
homeostasis. Using Fura-2 AM method to estimate the intracellular Ca2+ from both
TG-NS and TG-WT myocytes, an irregular Ca2+ handling was also observed in TG-NS
myocytes which had longer recovery time when pacing at 0.5 Hz. Besides, the time to 90%
recovery (T90r) of caffeine-induced Ca2+ transient signals was also longer in TG-NS
65
cells.
In order to link the abnormal Ca2+ transient signals in TG-NS myocytes to the
presence of gain-of-function mutant Na v 1.5, a late sodium blocker, ranolazine, was
applied to the isolated TG-NS myocytes. Ranolazine is a blocker of the late sodium
current caused by LQT3 mutation(Fredj et al 2006, Zaza et al 2008). The ranolazine
treatment was performed in both isolated TG-NSL3 (Figure 7) and TG-NSL12 (Figure 8)
myocytes. After 1 hour of ranolazine treatment, the Ca2+ transient signal has significant
reduction (P < 0.05) in recovery time (T90r) (Figure 7E and Figure 8E) from caffeine
effect with little or no changes in the transient amplitudes. Before ranolazine treatment,
the myocytes in both TG-NSL3 (Figure 7A) and TG-NSL12 (Figure 8A) did not respond
to the stimulation and the interval between each signal was irregular. However, in the
presence of ranolazine, most isolated myocytes can respond to the stimulation (Figure 7C
and Figure 8C) in the absence of caffeine. These results indicate that although ranolazine
cannot reduce the higher SR Ca2+ storage significantly in TG-NS myocytes, it can reduce
the time required for the removal of Ca2+ from the cytosol. This is consistent with the
previous results that the SR Ca2+ content is not higher in TG-NS myocytes compared
with TG-WT, but T90r is longer for TG-NS myocytes.
66
Figure 7. Ca2+ Transients Signals in Isolated TG-NSL3 Mice Myocytes (age=6-10
months, n=14). (A & B) Representative Ca2+ transient recording in isolated TG-NSL3
myocytes with (B) and without (A) 10 mM of caffeine. (C & D) Representative Ca2+
transient recording in isolated TG-NSL3 myocytes treated with 10 µM of ranolazine for 1
hour with (D) and without (C) 10 mM of caffeine. (E) Summary of data from all the
recordings. The time needed for recovery of caffeine-induced transient signal is
significantly shortened after 10 µM of ranolazine treatment. *, P < 0.05 and **, P < 0.01
indicates statistical significance.
67
Figure 8. Ca2+ Transients Signals in Isolated TG-NSL12 Mice Myocytes (age=6-10
months, n=24). (A & B) Representative Ca2+ transient recording in isolated TG-NSL12
myocytes with (B) and without (A) 10mM caffeine. (C & D) Representative Ca2+
transient recordings from isolated TG-NSL12 myocytes treated with 10 µM of ranolazine
for 1 hour with (D) and without (C) 10 mM of caffeine. (E) Summary of data from all the
recordings. The time needed for recovery of caffeine-induced transient signal is
significantly shortened after 10 µM of ranolazine treatment. *, P < 0.05 and **, P < 0.01
indicates statistical significance.
68
3.5 Discussion
Previously, our lab has demonstrated that the LQTS-causing mutation N1325S in
cardiac sodium channel gene SCN5A led to abnormal cytosolic Ca2+ transient signals in
isolated single TG-NS ventricular myocytes. Then we have performed a series of
experiments showing the intracellular [Na+] i increase and the dysfunction of Na+/Ca2+
exchanger (NCX). These results gave us a potential mechanism that the abnormal Ca2+
handling in TG-NS cells is caused by this SCN5A gain-of-function mutation via
dysfunction of NCX. To the best of our knowledge, this is the first study that provides a
direct link between a SCN5A mutation and abnormal intracellular Ca2+ transient signaling.
The comparison between TG-NSL12 and TG-WTL10 mice which express the wild-type
human SCN5A gene and have the a similar copy number of the transgene revealed that
only TG-NS myocytes showed this abnormal Ca2+ handling. Besides, another transgenic
mouse line which carries less copy numbers of this mutant transgene, TG-NSL3, also
exhibited irregularities in Ca2+ transient signals, albeit with less severity. All these
observations indicate that it is the mutation N1325S, rather than the overexpression of
SCN5A, that caused the Ca2+ transient abnormality.
Our lab has also studied the expression level of the important Ca2+ handling proteins
in TG-NS mouse hearts using Western blot analysis. The up-regulation of L-type Ca2+
channel protein, NCX and RyR2 in the presence of N1325S mutation suggests the
remodeling of TG-NS myocytes. The changes in the Ca2+ handling proteins weaken the
linkage between SCN5A mutation and abnormal Ca2+ handling we have set based on
intracellular [Na+] increase and NCX dysfunction. In order to further investigate the
relationship between this gain-of-function mutation and the abnormal Ca2+ handling in
69
the TG-NS myocytes, we introduced the LQTS drug, ranolazine, to block the late sodium
current produced by N1325S mutation. After 1 hour of 10 µM ranolazine treatment, the
abnormal Ca2+ transient signal in the isolated TG-NSL3 and TG-NSL12 ventricular
myocytes can be successfully rescued. This rescue effect further strengthened the direct
linkage between N1325S mutation and abnormal Ca2+ handling.
In the previous study of Ca2+ transient signals in TG-NS ventricular myocytes, the
most remarkable Ca2+ transient irregularity in TG-NS cells was in response to caffeine
application though the amplitudes of the caffeine-induced Ca2+ transients did not have
significant differences when compared to TG-WT cells. The prolongation of the recovery
rate was often characterized by oscillations or “re-excitations” in TG-NS cells, and this
characterization was also observed in HL-1 cells expressing N1325S mutant human
SCN5A gene. These results suggested that the Ca2+ abnormalities in TG-NS ventricular
myocytes may not stem from an “excess” of SR Ca2+ ion content but rather from a
weakened ability to recover the cytosolic Ca2+ concentration to baseline. According to
Ca2+ handling study in cardiac cells, the decline of intracellular Ca2+ ions during a
caffeine-induced SR-Ca2+ release is almost entirely due to the extrusion of the Ca2+ via
NCX (Altamirano et al 2006, Picht et al 2007). The interesting finding is that the
expression of NCX-1 is increased in TG-NS mouse hearts, which seems to be
contradictory to our findings in Ca2+ handling. Whereas, it is believed that NCX-1
expression level increase is a response to this chronic elevation of cytosolic Ca2+ in
TG-NS myocytes. In the presence of N1325S mutant SCN5A gene, both intracellular Na+
and Ca2+ increase. It is hypothesized that the excess of Na+ and Ca2+ forces an adaptive
response of the TG-NS cells in which the NCX is impossible to work in its proper way.
70
The lack of NCX current in TG-NS myocytes gives us a strong support to our hypothesis.
As a result, in order to compensate dysfunction of NCX, more NCX-1 will be expressed
in TG-NS myocytes.
Based on the result of our experiments, the ranolazine treatment to isolated
TG-NSL3 and TG-NSL12 ventricular myocytes can significantly reduce the recovery
time of caffeine-induced Ca2+ transients. It suggests that one hour, 10 µM ranolazine
treatment is enough to recover the NCX function. The excess Na+ entry caused by I Na,L is
reduced using this late sodium blocker ranolazine, and the application of ranolazine can
improve Ca2+ homeostasis.
In conclusion, this study has established the link between the existence of cytosolic
Ca2+ inhomogeneity in TG-NS ventricular murine myocytes and the presence of a
gain-of-function LQTS mutation N1325S in human cardiac sodium channel gene SCN5A,
which is known to generate late sodium current INa,L . The application of INa,L blocker
ranolazine is proved to successfully rescue the abnormal Ca2+ handling in TG-NS mouse
ventricular myocytes. This finding along with the intracellular Ca2+ abnormality in the
presence of SCN5A mutation N1325S may provide a molecular mechanism for
development of VT and DCM in both TG-NS mice and human patients with SCN5A
mutations.
71
CHAPTER IV
LATE SODIUM CURRENT PRODUCED BY GAIN-OF-FUNCTION LQTS
MUTATIONS IN CARDIAC SODIUM CHANNEL LEADS TO ABNORMAL
CALCIUM HANDLING IN CARDIAC CELLS
4.1 Abstract
N1325S mutation in the cardiac sodium channel gene SCN5A produces late sodium
currents (I Na,L ) and is associated with type-3 long QT syndrome (LQTS), ventricular
tachyarrhythmias (VT) and dilated cardiomyopathy (DCM) in both human and mice.
However, the underlying molecular mechanisms that originated from I Na,L need to be
further explored. In this study, I have successfully correlated the mutant SCN5A with
abnormal Ca2+ transient signals in transfected HL-1 cells. In HL-1 cells expressing
N1325S mutant human SCN5A, the caffeine-induced Ca2+ transient signal showed
significant prolonged recovery time from caffeine effects compared with cells expressing
wild type human sodium channel Na v 1.5. This result indicates the reduced ability to
72
re-uptake the cytosolic Ca2+ of SR or to remove extra intracellular Ca2+ out of the cell in
the presence of mutant Na v 1.5. Based on these observations, I hypothesized that INa,L
leads to cytosolic Ca2+abnormalities. In order to test my hypothesis, I treated the cells
with late sodium current blockers, ranolazine and mexiletine, and test if the abnormal
Ca2+ handling can be rescued. The experiment results confirmed that both blockers was
able to shorten the time required to remove the cytosolic Ca2+ from caffeine effects with
no significant reduction in peak Ca2+ response. In order to further test my hypothesis, I
then selected several other SCN5A mutations that can either cause LQTS or Brugada
syndrome for SCN5A mutation classification. In this experiment, I measured the Ca2+
transients and time of recovery from caffeine in these mutations. After measurements, the
results were compared in different types of Na v 1.5 mutations. The mutation classification
results further indicated the correlation between the late sodium current and the weakened
ability of the cell to reduce the cytosolic Ca2+. Taken together, this study first correlates
sodium mutations with Ca2+ handling in cardiac cells, and it can provide us important
insights into the pathogenic mechanisms of LQTS, VT and DCM.
4.2 Introduction
The transient opening of Na+ channels in cardiac cells gives rise to the inward
current responsible for the upstroke of the cardiac action potential (AP) and the initiation
of excitation-contraction coupling (Kiyosue and Arita 1989, Li et al 1997, Liu et al 1992).
The changes in cardiac AP are composed of 5 phases from 0 to 4 and the opening of
sodium channel contributes to rapid depolarization phase 0 and the plateau phase 2. After
opening, sodium channels enter an inactivated state. However, gain-of-function mutant
73
sodium channels will stay at the inactivated state during the plateau (phase 2) of the AP
when the wild type channels are usually closed. The abnormal prolonged inactivated state
of the sodium channels will allow the continuous flow of Na+ into the cell and results in
persistent late sodium current (I Na,L ). Accordingly, the presence of I Na,L slows down the
repolarization of the cell and therefore increases action potential duration (APD) (Liu et
al 1992).
It is known that under ischemia/reperfusion conditions, sodium channel activation
in cardiac cells is impaired, and that leads to INa,L . Consequently, I Na,L will cause the rise
of myocardial Na+ content (Barrington et al 1997, Williams et al 2007). The increase of
intracellular Na+ content can lead to increased Ca2+ via Na+/Ca2+ exchanger (NCX),
which is consistent with the findings that I Na,L leads to a Na+-dependent Ca2+ overload
(Fraser et al 2006, Pieske et al 2002, Song et al 2006, Wang et al 2008).
N1325S mutation in cardiac sodium channel SCN5A gene is a type-3 long QT
syndrome mutation and was originally identified in a 96-member family (Wang et al
1995a). Previously, our lab generated a transgenic mouse model that overexpresses
N1325S mutant human SCN5A in the heart (TG-NSL3 and TG-NSL12). According to the
prior results, I Na,L produced by this gain-of-function LQTS mutation results in prolonged
APDs in isolated ventricular myocytes, increased QTc on electrocardiograms and a high
incidence of both ventricular tachyarrhythmias (VT) and sudden death in both TG-NSL3
and TG-NSL12 mice (Tian et al 2004). Besides, Dr. Sandro Yong in our lab has found
that the Ca2+ transient signal is abnormal in isolated TG-NS myocytes due to the elevated
intracellular Na+ concentration and Na+/Ca2+ dysfunction. All these findings in TG-NS
mice demonstrate a potential mechanism that can explain the pathogenesis of this
74
mutation. However, there is one more problem to be solved before we come up with the
conclusion. In TG-NS myocytes, some important Ca2+ handling proteins, e.g., NCX,
RyR2 and L-type Ca2+ channel, are up-regulated in the presence of N1325S mutation
(Zhang et al 2011). The up-regulation of these proteins indicates the remodeling of adult
TG-NS myocytes caused by the mutant SCN5A transgene.
In order to overcome the adult myocytes remodeling problem, I introduced HL-1
cells, a cardiac muscle cell line derived from AT-1 mouse atrial cardiomyocyte tumor
lineage, in this study (Claycomb et al 1998). In HL-1 cells expressing N1325S mutant
SCN5A gene, the time required to remove extra cytosolic Ca2+ induced by caffeine is
significantly prolonged compared with cells expressing wild type SCN5A gene. The
further study of the effect for the blockage of I Na,L on abnormal Ca2+ handling provides us
more evidences that I Na,L correlates with Ca2+ handling abnormality in cardiac cells
expressing gain-of-function mutant Na v 1.5. The linkage between SCN5A mutation and
abnormal Ca2+ handling offers a potential mechanism for LQTS and cardiac arrhythmias.
4.3 Materials and Methods
4.3.1
HL-1 Cardiac Cell Culture
HL-1 cardiac cell was a gift from Dr. William Claycomb (Louisiana State
University, Medical Center, New Orleans, LA). According to the instruction from Dr.
Claycomb, the cells were cultured in 5% CO 2 at 37oC in Claycomb media
(Sigma-Aldrich, St. Louis, MO) supplemented with batch-specific 10% FBS (SigmaAldrich, St. Louis, MO), 100 U-100 µg/ml penicillin-streptomycin (medium room at
Lerner Research Institute, Cleveland Clinic Foundation, Cleveland, OH), 2 mM L-
75
glutamine (medium room at Lerner Research Institute, Cleveland Clinic Foundation,
Cleveland, OH), and 0.1 mM norepinephrine (Sigma-Aldrich, St. Louis, MO). Before
plating HL-1 cells, all the culture flasks and dishes were pre-treated for overnight with
fibronectin/gelatin coating buffer containing 0.02% gelatin (ThermoFisher Scientific,
Austin, TX) with 0.5% fibronectin (Life Technologies, Grand Island, NY).
4.3.2
Na+ Channel Cloning and Site-Directed Mutagenesis
Both wild type (WT) and mutant human cardiac SCN5A constructs were generated
as described in the previous study (Wan et al 2001). All the plasmids were prepared at the
same time using plasmid Maxi Prep (QIAGEN, Valencia, CA).
4.3.3
Transient Transfection of HL-1 Cell
HL-1 cardiac cells were transfected with either wild-type (WT) or mutant SCN5A
expression plasmids along with GFP as an indicator for transfected cells using the PolyJet
DNA In Vitro Transfection Reagent (SignaGen Laboratories, Gaithersburg, MD)
according to the manufacturer’s instructions with minor modifications to optimize the
transfection efficiency. For Ca2+ transient measurements, HL-1 cells were plated on 12
mm diameter glass coverslips (Life Technologies, Grand Island, NY) at a density of 3 ×
105 cells/ well in a 12-well culture plate (Sigma-Aldrich, St. Louis, MO). For each well,
the amount of plasmid DNA and PolyJet reagent used for transfection is 1 µg and 3 µL,
respectively. The transfected HL-1 cells were ready for Ca2+ transient measurements 48
hours after transfection.
76
4.3.4
Intracellular Ca2+ Transient Measurements
Before Ca2+ transient measurements, the transfected HL-1 cardiac cells were
incubated with 10 µM Fura-2 acetoxymethyl ester (Fura-2/AM, TEFlabs, Austin, TX) for
10 minutes at room temperature in the dark. The cells were then washed with the fresh
complemented Claycomb medium and incubated at 37oC with 5% CO2 for 20 min to
make sure that the acetoxymethyl group in Fura-2/AM is completely removed by cellular
esterase. The coverslips containing HL-1 cells were then transferred to a Bioptech
chamber on the microscopic stage at 28oC, and initial field stimulation was set at 0.5 Hz
and 5 ms duration during measurement. Ca2+ transient was measured simultaneously in
20-30 cells using excitation (340/380 nm) and emission (510 nm) spectra from the
spectofluorometer with a fixed test field. As an index of Ca2+ changes, the ratio of the
emitted fluorescence at the wavelength of 340nm over 380nm (F/F o ) was calculated.
Steady state Ca2+ transient signals were obtained at a pacing frequency of 0.25 Hz in the
absence of caffeine. The relative sarcoplasmic reticulum (SR) Ca2+ contents of HL-1
cardiac cells expressing both WT and mutant SCN5A were estimated by comparing the
amplitudes of Ca2+ transients before and after the application of 10mM caffeine, and the
time needed for 90% recovery from caffeine application was calculated. The stimulation
was stopped before 10 mM caffeine application. The rescue effects of late sodium current
blockers were tested in the presence of 10 µM ranolazine and 10 µM mexiletine after one
hour treatment at 37oC with 5% CO 2 .
4.3.5
Data Analysis
All the results were expressed as mean ± S.E.M., and the differences between
77
means were calculated by student’s paired t-test. A value of P < 0.05 was considered as
statistically significant.
4.4 Results
4.4.1
N1325S Mutant Sodium Channels Lead to Irregular Ca2+ Handling in HL-1
Cells
In order to evaluate the effects of gain-of-function mutation N1325S in SCN5A
gene to the cardiac Ca2+ handling, the HL-1 cells were transfected with wild-type (WT)
or N1325S mutant SCN5A and used to assess the cytosolic Ca2+ levels using the Fura-2
AM. In excited HL-1 cells expressing the N1325S mutation, I found the premature or
re-excited Ca2+ transient signals, which did not respond accordingly to the stimulation at
a 0.25 Hz frequency. This abnormal Ca2+ transient signals indicate that the cytosolic Ca2+
in HL-1 cells expressing N1325S mutant SCN5A has difficulties in returning to baseline
under 0.25 Hz stimulation frequency.
In order to further estimate the SR Ca2+ storage in mutant cells, I acutely treated the
transfected cells with 10 mM caffeine. Caffeine is commonly used to estimate the content
of the SR Ca+ storage, Ca2+ release, and the rate of Ca2+ extrusion from the cells (Picht et
al 2006). In this experiment, I observed a significant increase in the amplitude and
duration of Ca2+ release after adding caffeine (10 mM) to the cells (Figure 9C) in HL-1
cells expressing N1325S mutant SCN5A. The amplitude of the Ca2+ transient in response
to caffeine is often an index of the SR Ca2+ content during diastole (Lewartowski and
Zdanowski 1990). In HL-1 cells expressing N1325S mutant SCN5A, the higher Ca2+
transient amplitudes comparing with HL-1 cells expressing WT SCN5A indicate a higher
78
SR Ca2+ content during diastole (F/F o value: 0.53 ± 0.03 in N1325S and 0.35 ± 0.03 in
WT, P < 0.05). At the same time, the time required for Ca2+ transients returning to
baseline in response to caffeine is higher in cells expressing N1325S mutation than in
control cells (20.5 ± 1.9 s in N1325S and 14.7 ± 0.8 s in WT, P < 0.05). These results
indicate that compared with cells expressing WT SCN5A, HL-1 cells with the N1325S
mutation have higher initial SR Ca2+ content, and slower removal of Ca2+ from the
cytosol compared with WT cells. Besides, in some HL-1 cells expressing the N1325S
mutation, there are a second caffeine-induced Ca2+ responses (Figure 9D). This
phenomenon can be explained by re-excitation by elevated cytosolic Ca2+ concentrations
of the cells during repolarization. The sustained high cytosolic Ca2+ causes the release of
SR Ca2+ storage more than once during one excitation-contraction coupling.
79
N1325S
-Caffeine
A
N1325S
+Caffeine
F0/F=0.1
F0/F=0.1
10s
10s
WT
+Caffeine
WT
-Caffeine
B
F0/F=0.1
F0/F=0.1
10s
10s
C
Amplitude before caffeine
0.6
0.30
Amplitude after caffeine
*
0.5
0.25
0.4
0.20
0.3
0.15
0.2
0.10
0.1
0.05
0
0
90% recovery time T90 before caffeine
25
3.2
2.8
2.4
2.0
1.6
1.2
0.8
0.4
0
90% recovery time T90 after caffeine
*
20
15
10
5
0
D
Ca2+ oscillations or “re-excitations”
N1325S
+Caffeine
Figure 9. Ca2+ Transients Signals in Transfected HL-1 Cells. (A) Representative Ca2+
transient recordings in HL-1 cells expressing N1325S mutation in SCN5A gene with
(right) and without (left) 10 mM of caffeine. (B) Representative Ca2+ transient recordings
in HL-1 cells expressing wild type SCN5A gene with (right) and without (left) 10 mM
caffeine. (C) Summary of data from all the recordings. The caffeine-induced peak Ca2+
response is higher in N1325S HL-1 cells (upper right). The time needed for recovery
from caffeine effect is also significantly prolonged in HL-1 cells expressing N1325S
mutant SCN5A gene. (D) Representative Ca2+ transient signals with re-excitation after
adding caffeine. *, P < 0.05 indicates statistical significance.
80
4.4.2
Ranolazine Can Reduce the SR Ca2+ Content and Strengthen the Ability of
Removing Extra Cytosolic Ca2+ in HL-1 Cells Expressing N1325S Mutant
SCN5A
N1325S mutation is a gain-of-function mutation in SCN5A that can produce
persistent late sodium currents (Tian et al 2004). In the isolated myocytes from transgenic
mice carrying N1325S mutation (TG-NS), the intracellular Na+ concentration is higher
compared with non-transgenic (NTG) and transgenic wild type (TG-WT) mice. During
the excitation-contraction coupling in the cardiac cells, Ca2+ ions are extruded from the
cells mainly via the Na+/Ca2+ exchanger in order to reduce the cytosolic Ca2+ after
diastole (Bers 2002). In our TG-NS mouse myocytes, the lack of Ni+-sensitive current
indicated that Na+/Ca2+ exchanger is dysfunctional (Yong S, unpublished data). These
two findings in our TG-NS mice provided a direct link between a SCN5A mutation and
the abnormal intracellular Ca2+ signaling. However, the up-regulation of some Ca2+
handling proteins such as NCX, RyR2 and L-type Ca2+ channel indicated the remodeling
of the cardiac cells in adult TG-NS mice (Zhang et al 2011). The lack of mature cardiac
cell phenotype in neonatal cardiomyocytes limits the use of isolated neonatal myocytes in
TG-NS mice. In this study, the introduction of HL-1 cells overcame the remodeling
problem in adult TG-NS mice.
To correlate the effects of late sodium current and abnormal Ca2+ handling, the
cells were treated with 10 µM of ranolazine (Figure 10) to block the late sodium currents.
Ranolazine is a investigative anti-arrhythmic drug in phase III trials. It has been approved
that ranolazine can successfully block the late sodium currents caused by LQT3
mutations in the SCN5A gene (Fredj et al 2006, Zaza et al 2008). After 1 hour of
81
ranolazine treatment, I observed that the Ca2+ transient signal showed significant
reduction from 20.5 ± 1.9 s to 12.9 ± 1.7 s (P < 0.05) in recovery time (T90) (Figure 10F)
from caffeine effect with little or no changes in the transient amplitudes (Figure 10D).
Besides, in the presence of ranolazine, most abnormal transient signals can also be
rescued to normal and can respond to the stimulation accordingly (Figure 10A and 10B).
These results indicate that although ranolazine cannot reduce the higher SR Ca2+ storage
significantly in N1325S mutant HL-1 cells (Figure 10D), it can reduce the time required
for the removal of Ca2+ from the cytosol (Figure 10F). Therefore, according to my results,
the treatment of ranolazine itself is not enough to reduce the SR Ca2+ content significantly.
However, the shortening of time that is required for the removal of Ca2+ from cytosol can
explain the potential mechanism why ranolazine is an effective treatment for LQT3
patients. Consequently, after successful blockage of late sodium currents by ranolazine,
the elevated intracellular Na+ concentration is reduced, and the dysfunction of NCX is
rescued. Note that during the excitation-contraction coupling in the cardiac cells, Ca2+
ions are extruded from the cells mainly via the Na+/Ca2+ exchanger in order to reduce the
cytosolic Ca2+ after diastole (Bers 2002). The re-gaining function of NCX increases the
ability of the cell to extrude cytosolic Ca2+.
Furthermore, ranolazine treatment was not able to decrease the caffeine-induced
peak Ca2+ responses (Figure 10D). One of the excitation-contraction coupling model
proposed that the SR Ca2+ is stored in two compartments: an uptake compartment and a
release compartment (Euler 1999). Thus, the caffeine-induced peak Ca2+ response may
not be determined by the total SR Ca2+ content, but the Ca2+ content available to be
released in the release compartment. The ranolazine treatment for one hour is not enough
82
to break the “balance” between uptake compartment and release compartment Ca2+
content in SR, thus it would not change the peak Ca2+ response in the presence of 10 mM
caffeine.
83
-Ranolazine
+Caffeine
-Ranolazine
-Caffeine
A
F0/F=0.1
F0/F=0.1
10s
10s
+Ranolazine
+Caffeine
+Ranolazine
-Caffeine
B
F0/F=0.1
F0/F=0.1
10s
10s
C
0.30
F/F0
Amplitude before caffeine
D
0.25
F/F0 0.5
0.20
0.4
0.15
0.3
0.10
0.2
0.05
0.1
Sec 3.2
Amplitude after caffeine
0
0
E
0.6
90% recovery time T90 before caffeine
F
2.8
25
90% recovery time T90 after caffeine
20
2.4
2.0
15
*
1.6
10
1.2
0.8
5
0.4
0
0
HL-1 cells carry N1325S mutation in SCN5A without Ranolazine treatment
HL-1 cells carry N1325S mutation in SCN5A with 10uM Ranolazine treatment for 1 hour
Figure 10. Effect of Ranolazine Treatments in HL-1 Cells Expressing the N1325S
Mutation. (A) Representative Ca2+ transient recordings in HL-1 cells carrying the
N1325S mutation before (left) and after (right) caffeine. (B) Reprensentative Ca2+
transient recording in HL-1 cells carrying the N1325S mutation in the presence of 10 µM
of ranolazine treatment before (left) and after (right) caffeine. (C-F) Summary of data
from all the recordings. The time needed for the removal of cytosolic Ca2+ after caffeine
is shortened after ranolazine treatment. *, P < 0.05 indicates statistical significance.
84
4.4.3
Mexiletine Can Reduce the SR Ca2+ Content in HL-1 Cells Expressing
N1325S Mutant SCN5A
In this study, another late sodium current blocker mexiletine was also tested in
N1325S mutant HL-1 cells. Mexiletine can abbreviate the QT interval in LQT3 patients
(Sicouri et al 1997). Our lab previously reported that mexiletine was able to successfully
block the late sodium current in isolated TG-NS myocytes (Tian et al 2004). Mexiletine
can also suppress ventricular fibrillation (VF) and ventricular tachycardia (VT) in TG-RQ
mice (Tian et al 2004). In this study, I identified that similar to ranolazine, in the presence
of 10 µM of mexiletine, the recovery time from caffeine is significantly reduced by ~50%
from more than 20 seconds to 10.6 ± 1.4 s (P < 0.05) (Figure 11F). And also similar to
ranolazine, mexiletine treatment is also not able to reduce the caffeine-induced Ca2+ peak
response (Figure 11E). And the cells can respond to the stimulation better after treatment
(Figure 11A and 11B).
85
-Mexiletine
+Caffeine
-Mexiletine
-Caffeine
A
F0/F=0.1
F0/F=0.1
10s
10s
B
+Mexiletine
+Caffeine
+Mexiletine
-Caffeine
F0/F=0.1
F0/F=0.1
10s
C
0.3
10s
D
Amplitude before caffeine
1.4
0.25
1.2
0.2
1
0.15
0.8
0.1
0.6
0.4
0.05
0.2
0
E
1.6
0.6
90% recovery time T90 before caffeine
0
Amplitude after caffeine
F
0.5
25
90% recovery time T90 after caffeine
20
0.4
15
0.3
*
10
0.2
5
0.1
0
0
HL-1 cells carry N1325S mutation in SCN5A without Mexiletine treatment
HL-1 cells carry N1325S mutation in SCN5A with 10uM Mexiletine treatment
Figure 11. Effect of Mexiletine Treatments in HL-1 Cells Expressing the N1325S
Mutation. (A) Representative Ca2+ transient recordings in HL-1 cells carrying the
N1325S mutation before (left) and after (right) caffeine. (B) Reprensentative Ca2+
transient recording in HL-1 cells carrying the N1325S mutation in the presence of 10 µM
of mexiletine treatment before (left) and after (right) caffeine. (C-F) Summary of data
from all the recordings. The time needed for the removal of cytosolic Ca2+ after caffeine
is shortened after mexiletine treatment. *, P < 0.05 indicates statistical significance.
86
4.4.4
Classification of Different SCN5A Mutations
Since LQTS mutation N1325S showed these Ca2+ transient signal specific features
in both isolated myocytes and transfected HL-1 cells, I was wondering if other
gain-of-function mutations in SCN5A will display the same phenotype or it is just
N1325S specific phenomenon. In the next series of experiments, I have selected other
two LQTS mutations in SCN5A gene, E1784K and ΔKPQ, which also produce late
sodium currents like N1325S mutation (Dumaine et al 1996, Wei et al 1999). At the same
time, I selected two Brugada syndrome mutations, A1924T and L567Q, which do not
produce the late sodium current (Rook et al 1999, Wan et al 2001). In this experiment, I
only evaluated the effects from caffeine effects to estimate the content of SR Ca2+ storage
and the recovery time required to remove extra cytosolic Ca2+ (Figure 12). After
statistical analysis, I found that both E1784K and ΔKPQ mutations can significantly
prolong the recovery time from caffeine, and the prolongation of E1784K mutation is
almost 10 times as much as wild type sodium channel (Figure 12C). The caffeine-induced
peak Ca2+ response in N1325S, ΔKPQ, E1784K and A1924T are all significantly higher
than HL-1 cells transfected with WT SCN5A (Figure 12B). According to my results, the
caffeine-induced peak Ca2+ response is not correlated with the presence of LQTS
mutations (Figure 12B). The reason behind this could be explained by the two SR
compartments theory(Euler 1999). During excitation-contraction coupling, the cells
expressing LQTS mutant SCN5A experience a spontaneous Ca2+ oscillation, and no
sufficient time for the cells to transfer the uptake compartment Ca 2+ to release
compartment in SR. As a result, the Ca2+ content in release compartment is limited and
not be able to reflect the total SR Ca2+ content. This theory about SR Ca2+ storage can
87
180
160
140
120
100
80
60
40
20
0
Peak Ca2+ response (%)
B
WT
ΔKPQ
N1325S
**
**
E1784K
C
**
90% Transient recovery (sec)
A
180
160
140
120
100
80
60
40
20
0
A1924T
L567Q
**
*
**
Figure 12. Effects of Different SCN5A Mutations on Ca2+ Transients. (A) Representative
Ca2+ signal response induced by 10 mM of caffeine. (B) The peak Ca2+ response in
different mutations related to wild type SCN5A (100%). Three mutations, N1325S, ΔKPQ
and A1924T have higher peak Ca2+ response compared with wild type sodium channels.
(C) Time required for 90% recovery in the intracellular Ca2+ from caffeine effect. Three
mutations, N1325S, ΔKPQ and E1784K which are all LQTS mutations producing late
sodium currents need significantly longer 90% transient recovery time. *, P < 0.05 and
**, P < 0.01 indicate statistical significance.
88
explain why the caffeine-induced peak Ca2+ response is not always higher in LQTS
mutant cells. Except for Na2+ content, the amount of Ca2+ in SR release compartment is
affected by other factors or it is a result of complex factors.
4.4.5 Ranolazine Effect in Various SCN5A Mutations
In order to further test the relationship between different types of SCN5A mutations
and Ca2+ handling, the ranolazine effect in these mutations was also tested (Figure 13).
After one hour, 10 µM of ranolazine treatment, the Ca2+ transient signals were collected
in HL-1 cells transfected with different SCN5A mutations and wild-type SCN5A. As I
expected, the 90% transient recovery time from caffeine was significantly reduced in the
mutations (N1325S, ΔKPQ and E1784K) that can produce certain amount of late sodium
currents (Figure 13C). The shortening of recovery time from caffeine is especially
dramatic in E1784K mutation, from 132.67 seconds (n=8) to an average of 13.877
seconds (n=10). At the same time, the peak Ca2+ response from caffeine in E1784K
mutation is more than 300% compared with wild-type SCN5A (Figure 13B). This
significantly higher peak Ca2+ response from caffeine in E1784K mutation indicates an
increase of release Ca2+ content in SR after ranolazine treatment. One of the Brugada
mutations, L567Q, also showed significantly increased peak Ca2+ response in the
presence of caffeine, but the mechanism behind this is not clear and need to be further
explored (Figure 13B).
89
Figure 13. Effect of Ranolazine in Various SCN5A Mutations. (A) Representative Ca2+
signal induced by 10 mM caffeine after 1 hour, 10 µM ranolazine treatment. (B) The peak
Ca2+ response in different mutations related to wild type SCN5A (100%) after ranolazine
treatment. One of the LQTS mutations, E1784K, has significantly increased response
compared with WT. One of the Brugada mutations, L567Q, has dramatically stronger
caffeine response compared with WT. (C) Time required for 90% recovery in the
intracellular Ca2+ from caffeine effect. After one hour, 10 µM ranolazine treatment, all the
LQTS mutations studied in this experiment showed no significant differences compared
with WT. *, P < 0.05 and **, P < 0.01 indicate statistical significance.
90
4.5 Discussion
In this study, I introduced transfected HL-1 cells for Ca2+ handling study related to
Na v 1.5 mutations for the first time. The results in this study demonstrate that
gain-of-function LQTS causing mutation, N1325S, in cardiac sodium channel gene
SCN5A leads to abnormal cytosolic Ca2+ transient signal in HL-1 cells. This Ca2+
handling study correlates the sodium mutations with abnormal intracellular Ca2+ in the
cardiac cells. The electrophysiological study of transgenic mice is time consuming, and
the isolation for adult mice cardiomyocytes depends on the quality of the collagenase
from the commercial companies. The introduction of HL-1 cells will solve the adult
myocytes remodeling problems because HL-1 cells are easy to be transfected and can
retain the adult cardiomyocyte phenotype in culture.
In this study, I found that the caffeine-induced peak Ca2+ transients in HL-1 cells
expressing N1325S mutant SCN5A were increased, and the recovery time from caffeine is
significantly prolonged in mutant cells. The presence of Ca2+ oscillations or
“re-excitations” in N1325S HL-1 cells suggests that the abnormal Ca2+ handling may not
stem from the higher content of SR Ca2+ content, but from a reduced ability to remove
the cytosolic Ca2+ back to SR or out of the cells. Therefore, the prolonged reduction of
cytosolic Ca2+ can cause another Ca2+ induce Ca2+ release (CICR) and give a second peak
when adding caffeine. This result indicates the linkage between SCN5A mutation and
abnormal Ca2+, the following series of experiments I had designed would further
strengthen the correlation between these two.
In order to investigate whether late sodium currents produced by N1325S mutation
play a critical role in abnormal intracellular Ca2+ transient signals, I treated the cells with
91
late sodium current blockers, ranolazine and mexiletine. Following the treatment of 10
µM of ranolazine or mexiletine, the recovery time from caffeine effects is significantly
shortened with no change in amplitudes. This suggested that a reduction of excessive Na+
entry caused by late sodium currents will reduce the intracellular Na+ concentration and
quicken the removal of cytosolic Ca2+ back to SR or out of the cells. The comparison of
different SCN5A mutations on Ca2+ handling indicates that the SR Ca2+ content was not
necessarily affected by late sodium currents because not only did N1325S, ΔKPQ and
E1784K mutations increase the caffeine-induced peak Ca2+ response, but A1924T
mutation also increased the peak Ca2+ response. A1924T mutation is a Brugada mutation
causing negative shift of inactivation curve, I Na increase and a larger action potential
overshoot (Rook et al 1999). The overall impact on intracellular Na+ concentration is
unknown. However, according to my results, the ability of removing extra cytosolic Ca2+
is correlated to the presence of late sodium currents produced by SCN5A mutations
because only mutations that produce I Na,L have significantly prolonged recovery time
from caffeine.
Since the first mutation in SCN5A related to cardiac arrhythmia was identified in
1995 (Wang et al 1995b), various mutations in this gene have been found and associated
with numerous arrhythmia syndromes, such as LQT3, Brugada syndrome, progressive
cardiac conduction defect (PCCD), sick sinus node syndrome (SSS), atrial fibrillation
(AF) and DCM. The molecular pathology of these mutations is still not fully understood.
Ca2+ is a crucial element in all kinds of organisms, and it plays an important role in
numerous biological functions. It is known that the intracellular free Ca2+ concentration
(~10-7 M) is much lower than the extracellular Ca2+. This high Ca2+ gradient provides
92
abundant Ca2+ available to be imported into cells, where these Ca2+ ions can function as
second messengers (Chin and Means 2000). Thus, the abnormal Ca2+ handling in some
SCN5A mutations could play an important role in the development of cardiac arrhythmia.
In conclusion, this study has established the correlation between gain-of-function
sodium mutations and abnormal cytosolic Ca2+ handling in cardiac cells. I believe that the
late sodium current is the cause for Ca2+ inhomogeneity caused by the N1325S mutation.
I further showed that in the presence of late sodium current blocker, ranolazine or
mexiletine, the prolonged recovery time from caffeine can be shortened. The advantage
of this study over the isolated TG-NS adult myocytes is the introduction of HL-1 cells.
HL-1 cells diminished the remodeling responses such as a higher apoptosis rate and
up-regulations of Ca2+ handling proteins in aging TG-NS mouse hearts. The finding that
abnormal intracellular Ca2+ transient signals are associated with gain-of-function LQTS
SCN5A mutation may provide a molecular mechanism for the development of VT and
DCM.
93
BIBLIOGRAPHY
Abbott GW, Sesti F, Splawski I, Buck ME, Lehmann MH, Timothy KW et al (1999).
MiRP1 forms IKr potassium channels with HERG and is associated with cardiac
arrhythmia. Cell 97: 175-187.
Abriel H, Kamynina E, Horisberger JD, Staub O (2000). Regulation of the cardiac
voltage-gated Na+ channel (H1) by the ubiquitin-protein ligase Nedd4. FEBS Lett 466:
377-380.
Abriel H, Kass RS (2005). Regulation of the voltage-gated cardiac sodium channel
Nav1.5 by interacting proteins. Trends Cardiovasc Med 15: 35-40.
Abriel H (2007). Cardiac sodium channel Nav1.5 and its associated proteins. Arch Mal
Coeur Vaiss 100: 787-793.
Ahern CA, Zhang JF, Wookalis MJ, Horn R (2005). Modulation of the cardiac sodium
channel NaV1.5 by Fyn, a Src family tyrosine kinase. Circ Res 96: 991-998.
Aitken A (2006). 14-3-3 proteins: a historic overview. Semin Cancer Biol 16: 162-172.
Akai J, Makita N, Sakurada H, Shirai N, Ueda K, Kitabatake A et al (2000). A novel
SCN5A mutation associated with idiopathic ventricular fibrillation without typical ECG
findings of Brugada syndrome. FEBS Lett 479: 29-34.
94
Allouis M, Le Bouffant F, Wilders R, Péroz D, Schott JJ, Noireaud J et al (2006). 14-3-3
is a regulator of the cardiac voltage-gated sodium channel Nav1.5. Circ Res 98:
1538-1546.
Altamirano J, Li Y, DeSantiago J, Piacentino V, Houser SR, Bers DM (2006). The
inotropic effect of cardioactive glycosides in ventricular myocytes requires Na+-Ca2+
exchanger function. J Physiol 575: 845-854.
Andavan GS, Lemmens-Gruber R (2011). Voltage-gated sodium channels: mutations,
channelopathies and targets. Curr Med Chem 18: 377-397.
Antzelevitch C, Pollevick GD, Cordeiro JM, Casis O, Sanguinetti MC, Aizawa Y et al
(2007). Loss-of-function mutations in the cardiac calcium channel underlie a new clinical
entity characterized by ST-segment elevation, short QT intervals, and sudden cardiac
death. Circulation 115: 442-449.
Antzelevitch C, Burashnikov A, Sicouri S, Belardinelli L (2011). Electrophysiologic basis
for the antiarrhythmic actions of ranolazine. Heart Rhythm 8: 1281-1290.
Asahi M, Nakayama H, Tada M, Otsu K (2003). Regulation of sarco(endo)plasmic
reticulum Ca2+ adenosine triphosphatase by phospholamban and sarcolipin: implication
for cardiac hypertrophy and failure. Trends Cardiovasc Med 13: 152-157.
95
Baba S, Dun W, Boyden PA (2004). Can PKA activators rescue Na+ channel function in
epicardial border zone cells that survive in the infarcted canine heart. Cardiovasc Res 64:
260-267.
Balijepalli RC, Kamp TJ (2008). Caveolae, ion channels and cardiac arrhythmias. Prog
Biophys Mol Biol 98: 149-160.
Balser JR, Bennett PB, Hondeghem LM, Roden DM (1991). Suppression of
time-dependent outward current in guinea pig ventricular myocytes: actions of quinidine
and amiodarone. Circ Res 69: 519-529.
Baroudi G, Carbonneau E, Pouliot V, Chahine M (2000). SCN5A mutation (T1620M)
causing Brugada syndrome exhibits different phenotypes when expressed in Xenopus
oocytes and mammalian cells. FEBS Lett 467: 12-16.
Barrington PL, Martin RL, Zhang K (1997). Slowly inactivating sodium currents are
reduced by exposure to oxidative stress. J Mol Cell Cardiol 29: 3251-3265.
Bassani JW, Bassani RA, Bers DM (1994). Relaxation in rabbit and rat cardiac cells:
species-dependent differences in cellular mechanisms. J Physiol 476: 279-293.
Baulac S, Gourfinkel-An I, Picard F, Rosenberg-Bourgin M, Prud'homme JF, Baulac M et
96
al (1999). A second locus for familial generalized epilepsy with febrile seizures plus
maps to chromosome 2q21-q33. Am J Hum Genet 65: 1078-1085.
Bazett HC (1920). An analysis of the time relations of electrocardiograms. Heart 7:
353-370.
Beggs AH (1997). Dystrophinopathy, the expanding phenotype. Dystrophin abnormalities
in X-linked dilated cardiomyopathy. Circulation 95: 2344-2347.
Bender AC, Morse RP, Scott RC, Holmes GL, Lenck-Santini PP (2012). SCN1A
mutations in Dravet syndrome: Impact of interneuron dysfunction on neural networks and
cognitive outcome. Epilepsy Behavior: E&B 23: 177-186.
Bennett PB, Yazawa K, Makita N, George AL, Jr. (1995). Molecular mechanism for an
inherited cardiac arrhythmia. Nature 375: 683-685.
Benson DW, Wang DW, Dyment M, Knilans TK, Fish FA, Strieper MJ et al (2003).
Congenital sick sinus syndrome caused by recessive mutations in the cardiac sodium
channel gene (SCN5A). J Clin Invest 112: 1019-1028.
Bers DM (2002). Cardiac excitation-contraction coupling. Nature 415: 198-205.
Beyder A, Strege PR, Reyes S, Bernard CE, Terzic A, Makielski J et al (2012).
97
Ranolazine decreases mechanosensitivity of the voltage-gated sodium ion channel
NaV1.5: a novel mechanism of drug action. Circulation 125: 2698-2706.
Bezzina C, Veldkamp MW, van Den Berg MP, Postma AV, Rook MB, Viersma JW et al
(1999). A single Na(+) channel mutation causing both long-QT and Brugada syndromes.
Circ Res 85: 1206-1213.
Bezzina CR, Tan HL (2002). Pharmacological rescue of mutant ion channels. Cardiovasc
Res 55: 229-232.
Bezzina CR, Rook MB, Groenewegen WA, Herfst LJ, van der Wal AC, Lam J et al
(2003). Compound heterozygosity for mutations (W156X and R225W) in SCN5A
associated with severe cardiac conduction disturbances and degenerative changes in the
conduction system. Circ Res 92: 159-168.
Binder WH, Barragan V, Menger FM (2003). Domains and rafts in lipid membranes.
Angew Chem Int Ed Engl 42: 5802-5827.
Blangy H, Sadoul N, Coutelour JM, Rebmann JP, Joseph M, Scherrer C et al (2005).
Prevalence of Brugada syndrome among 35,309 inhabitants of Lorraine screened at a
preventive medicine centre. Arch Mal Coeur Vaiss 98: 175-180.
Blaustein MP, Lederer WJ (1999). Sodium/calcium exchange: its physiological
98
implications. Physiol Rev 79: 763-854.
Braun AP, Schulman H (1995). The multifunctional calcium/calmodulin-dependent
protein kinase: from form to function. Annu Rev Physiol 57: 417-445.
Brink PA, Ferreira A, Moolman JC, Weymar HW, van der Merwe PL, Corfield VA (1995).
Gene for progressive familial heart block type I maps to chromosome 19q13. Circulation
91: 1633-1640.
Brodsky MA, Chun JG, Podrid PJ, Douban S, Allen BJ, Cygan R (1996). Regional
attitudes of generalists, specialists, and subspecialists about management of atrial
fibrillation. Arch Intern Med 156: 2553-2562.
Brown DA, London E (1998). Structure and origin of ordered lipid domains in biological
membranes. J Membr Biol 164: 103-114.
Brugada J, Brugada R, Antzelevitch C, Towbin J, Nademanee K, Brugada P (2002).
Long-term follow-up of individuals with the electrocardiographic pattern of right
bundle-branch block and ST-segment elevation in precordial leads V1 to V3. Circulation
105: 73-78.
Brugada P, Wellens HJ (1988). Arrhythmogenesis of antiarrhythmic drugs. Am J Cardiol
61: 1108-1111.
99
Brugada P, Brugada J (1992). Right bundle branch block, persistent ST segment elevation
and sudden cardiac death: a distinct clinical and electrocardiographic syndrome. A
multicenter report. J Am Coll Cardiol 20: 1391-1396.
Burgess DL, Kohrman DC, Galt J, Plummer NW, Jones JM, Spear B et al (1995).
Mutation of a new sodium channel gene, Scn8a, in the mouse mutant 'motor endplate
disease'. Nat Genet 10: 461-465.
Camacho JA, Hensellek S, Rougier JS, Blechschmidt S, Abriel H, Benndorf K et al
(2006). Modulation of Nav1.5 channel function by an alternatively spliced sequence in
the DII/DIII linker region. J Biol Chem 281: 9498-9506.
Cantrell AR, Smith RD, Goldin AL, Scheuer T, Catterall WA (1997). Dopaminergic
modulation of sodium current in hippocampal neurons via cAMP-dependent
phosphorylation of specific sites in the sodium channel alpha subunit. J Neurosci 17:
7330-7338.
Carmeliet E (1993). Use-dependent block of the delayed K+ current in rabbit ventricular
myocytes. Cardiovasc Drugs Ther 7: 599-604.
Casini S, Tan HL, Bhuiyan ZA, Bezzina CR, Barnett P, Cerbai E et al (2007).
Characterization of a novel SCN5A mutation associated with Brugada syndrome reveals
100
involvement of DIIIS4-S5 linker in slow inactivation. Cardiovasc Res 76: 418-429.
Chaitman BR, Pepine CJ, Parker JO, Skopal J, Chumakova G, Kuch J et al (2004).
Effects of ranolazine with atenolol, amlodipine, or diltiazem on exercise tolerance and
angina frequency in patients with severe chronic angina: a randomized controlled trial.
JAMA 291: 309-316.
Chambers JC, Zhao J, Terracciano CM, Bezzina CR, Zhang W, Kaba R et al (2010).
Genetic variation in SCN10A influences cardiac conduction. Nat Genet 42: 149-152.
Chang CC, Acharfi S, Wu MH, Chiang FT, Wang JK, Sung TC et al (2004). A novel
SCN5A mutation manifests as a malignant form of long QT syndrome with perinatal
onset of tachycardia/bradycardia. Cardiovasc Res 64: 268-278.
Chauhan VS, Tuvia S, Buhusi M, Bennett V, Grant AO (2000). Abnormal cardiac Na(+)
channel properties and QT heart rate adaptation in neonatal ankyrin(B) knockout mice.
Circ Res 86: 441-447.
Chen J, Chien KR (1999). Complexity in simplicity: monogenic disorders and complex
cardiomyopathies. J Clin Invest 103: 1483-1485.
Chen L, Marquardt ML, Tester DJ, Sampson KJ, Ackerman MJ, Kass RS (2007).
Mutation of an A-kinase-anchoring protein causes long-QT syndrome. Proc Natl Acad
101
Sci U S A 104: 20990-20995.
Chen Q, Kirsch GE, Zhang D, Brugada R, Brugada J, Brugada P et al (1998). Genetic
basis and molecular mechanism for idiopathic ventricular fibrillation. Nature 392:
293-296.
Chen S, Chung MK, Martin D, Rozich R, Tchou PJ, Wang Q (2002). SNP S1103Y in the
cardiac sodium channel gene SCN5A is associated with cardiac arrhythmias and sudden
death in a white family. J Med Genet 39: 913-915.
Chen T, Inoue M, Sheets MF (2005). Reduced voltage dependence of inactivation in the
SCN5A sodium channel mutation delF1617. Am J Physiol Heart Circ Physiol 288:
H2666-2676.
Chen YH, Xu SJ, Bendahhou S, Wang XL, Wang Y, Xu WY et al (2003). KCNQ1
gain-of-function mutation in familial atrial fibrillation. Science 299: 251-255.
Chin D, Means AR (2000). Calmodulin: a prototypical calcium sensor. Trends Cell Biol
10: 322-328.
Claycomb WC, Lanson NA, Jr., Stallworth BS, Egeland DB, Delcarpio JB, Bahinski A et
al (1998). HL-1 cells: a cardiac muscle cell line that contracts and retains phenotypic
characteristics of the adult cardiomyocyte. Proc Natl Acad Sci U S A 95: 2979-2984.
102
Collier JJ, White SM, Claycomb WC, Scott DK (2002). Insulin represses and chronic
hyperglycemia stimulates adrenomedullin gene expression in HL-1 cultured cardiac
myocytes. Diabetes 51.
Cordeiro JM, Barajas-Martinez H, Hong K, Burashnikov E, Pfeiffer R, Orsino AM et al
(2006). Compound heterozygous mutations P336L and I1660V in the human cardiac
sodium channel associated with the Brugada syndrome. Circulation 114: 2026-2033.
Couchonnal LF, Anderson ME (2008). The role of calmodulin kinase II in myocardial
physiology and disease. Physiology (Bethesda) 23: 151-159.
Cunha SR, Mohler PJ (2006). Cardiac ankyrins: Essential components for development
and maintenance of excitable membrane domains in heart. Cardiovasc Res 71: 22-29.
Curran ME, Splawski I, Timothy KW, Vincent GM, Green ED, Keating MT (1995). A
molecular basis for cardiac arrhythmia: HERG mutations cause long QT syndrome. Cell
80: 795-803.
Das S, Makino S, Melman YF, Shea MA, Goyal SB, Rosenzweig A et al (2009).
Mutation in the S3 segment of KCNQ1 results in familial lone atrial fibrillation. Heart
Rhythm 6: 1146-1153.
103
de Meeus A, Stephan E, Debrus S, Jean MK, Loiselet J, Weissenbach J et al (1995). An
isolated cardiac conduction disease maps to chromosome 19q. Circ Res 77: 735-740.
Deal KK, England SK, Tamkun MM (1996). Molecular physiology of cardiac potassium
channels. Physiol Rev 76: 49-67.
Deschênes I, Baroudi G, Berthet M, Barde I, Chalvidan T, Denjoy I et al (2000).
Electrophysiological characterization of SCN5A mutations causing long QT (E1784K)
and Brugada (R1512W and R1432G) syndromes. Cardiovasc Res 46: 55-65.
Dhar Malhotra J, Chen C, Rivolta I, Abriel H, Malhotra R, Mattei LN et al (2001).
Characterization of sodium channel alpha- and beta-subunits in rat and mouse cardiac
myocytes. Circulation 103: 1303-1310.
Dhein S, Müller A, Gerwin R, Klaus W (1993). Comparative study on the proarrhythmic
effects of some antiarrhythmic agents. Circulation 87: 617-630.
Dib-Hajj SD, Tyrrell L, Black JA, Waxman SG (1998). NaN, a novel voltage-gated Na
channel, is expressed preferentially in peripheral sensory neurons and down-regulated
after axotomy. Proc Natl Acad Sci U S A 95: 8963-8968.
Dib-Hajj SD, Tyrrell L, Escayg A, Wood PM, Meisler MH, Waxman SG (1999). Coding
sequence, genomic organization, and conserved chromosomal localization of the mouse
104
gene Scn11a encoding the sodium channel NaN. Genomics 59: 309-318.
Dobrzynski H, Boyett MR, Anderson RH (2007). New insights into pacemaker activity:
promoting understanding of sick sinus syndrome. Circulation 115: 1921-1932.
Dumaine R, Wang Q, Keating MT, Hartmann HA, Schwartz PJ, Brown AM et al (1996).
Multiple mechanisms of Na+ channel--linked long-QT syndrome. Circ Res 78: 916-924.
Echt DS, Liebson PR, Mitchell LB, Peters R, Obias-Manno D, Barker AH et al (1991).
Mortality and morbidity in patients receiving encainide, flecainide, or placebo. The
cardiac arrhythmia suppression trial. N Engl J Med 324: 781-788.
Euler DE (1999). Cardiac alternans: mechanisms and pathophysiological significance.
Cardiovasc Res 42: 583-590.
Fahmi AI, Patel M, Stevens EB, Fowden AL, John JE, 3rd., Lee K et al (2001). The
sodium channel beta-subunit SCN3b modulates the kinetics of SCN5a and is expressed
heterogeneously in sheep heart. J Physiol 537: 693-700.
Fatkin D, Otway R, Vandenberg JI (2007). Genes and atrial fibrillation: a new look at an
old problem. Circulation 116: 782-792.
Fink M, Noble PJ, Noble D (2011). Ca²⁺-induced delayed afterdepolarizations are
105
triggered by dyadic subspace Ca2²⁺ affirming that increasing SERCA reduces
aftercontractions. Am J Physiol Heart Circ Physiol 301: H921-935.
Fotia AB, Ekberg J, Adams DJ, Cook DI, Poronnik P, Kumar S (2004). Regulation of
neuronal voltage-gated sodium channels by the ubiquitin-protein ligases Nedd4 and
Nedd4-2. J Biol Chem 279: 28930-28935.
Fraser H, Belardinelli L, Wang L, Light PE, McVeigh JJ, Clanachan AS (2006).
Ranolazine decreases diastolic calcium accumulation caused by ATX-II or ischemia in rat
hearts. J Mol Cell Cardiol 41: 1031-1038.
Fredj S, Sampson KJ, Liu H, Kass RS (2006). Molecular basis of ranolazine block of
LQT-3 mutant sodium channels: evidence for site of action. Br J Pharmacol 148: 16-24.
Frohnwieser B, Chen LQ, Schreibmayer W, Kallen RG (1997). Modulation of the human
cardiac sodium channel alpha-subunit by cAMP-dependent protein kinase and the
responsible sequence domain. J Physiol 498: 309-318.
Garcia-Dorado D, Théroux P, Duran JM, Solares J, Alonso J, Sanz E et al (1992).
Selective inhibition of the contractile apparatus. A new approach to modification of
infarct size, infarct composition, and infarct geometry during coronary artery occlusion
and reperfusion. Circulation 85: 1160-1174.
106
Gellens ME, George AL, Jr., Chen LQ, Chahine M, Horn R, Barchi RL et al (1992).
Primary structure and functional expression of the human cardiac tetrodotoxin-insensitive
voltage-dependent sodium channel. Proc Natl Acad Sci U S A 89: 554-558.
George AL, Jr., Gellens ME, Kallen RG, Barchi RL (1990). Molecular cloning and
chromosomal location of two human muscle voltage-gated sodium channels. Soc
Neurosci Abst 16: 184.
George AL, Jr., Knops JF, Han J, Finley WH, Knittle TJ, Tamkun MM et al (1994).
Assignment of a human voltage-dependent sodium channel alpha-subunit gene (SCN6A)
to 2q21-q23. Genomics 19: 395-397.
George AL, Jr., Varkony TA, Drabkin HA, Han J, Knops JF, Finley WH et al (1995).
Assignment of the human heart tetrodotoxin-resistant voltage-gated Na+ channel
α-subunit gene (SCN5A) to band 3p21. Cytogenet Cell Genet 68: 67-70.
Glickman MH, Ciechanover A (2002). The ubiquitin-proteasome proteolytic pathway:
destruction for the sake of construction. Physiol Rev 82: 373-428.
Goetz R, Dover K, Laezza F, Shtraizent N, Huang X, Tchetchik D et al (2009). Crystal
structure of a fibroblast growth factor homologous factor (FHF) defines a conserved
surface on FHFs for binding and modulation of voltage-gated sodium channels. J Biol
Chem 284: 17883-17896.
107
Goldin AL (2008). Mechanisms of sodium channel inactivation. Curr Opin Neurobiol 13:
284-290.
Gollob MH, LJones D, Krahn AD, Gong LDXQ, Shao Q, Liu X et al (2006). Somatic
mutations in the connexin 40 gene (GJA5) in atrial fibrillation. N Engl J Med 354:
2677-2688.
Grace AA, Camm AJ (1998). Quinidine. N Engl J Med 338: 35-45.
Grant AO, Carboni MP, Neplioueva V, Starmer CF, Memmi M, Napolitano C et al (2002).
Long QT syndrome, Brugada syndrome, and conduction system disease are linked to a
single sodium channel mutation. J Clin Invest 110: 1201-1209.
Grant AO (2009). Cardiac ion channels. Circ Arrhythm Electrophysiol 2: 185-194.
Grynkiewicz G, Poenie M, Tsien RY (1985). A new generation of Ca2+ indicators with
greatly improved fluorescence properties. J Biol Chem 260: 3440-3450.
Hallaq H, Yang Z, Viswanathan PC, Fukuda K, Shen W, Wang DW et al (2006).
Quantitation of protein kinase A-mediated trafficking of cardiac sodium channels in
living cells. Cardiovasc Res 72: 250-261.
108
Han JA, Lu CM, Brown GB, Rado TA (1991). Direct amplification of a single dissected
chromosomal segment by polymerase chain reaction: a human brain sodium channel gene
is on chromosome 2q22-q23. Proc Natl Acad Sci U S A 88: 335-339.
Hanlon MR, Wallace BA (2002). Structure and function of voltage-dependent ion channel
regulatory beta subunits. Biochemistry 41: 2886-2894.
Hedley PL, Jørgensen P, Schlamowitz S, Moolman-Smook J, Kanters JK, Corfield VA et
al (2009). The genetic basis of Brugada syndrome: a mutation update. Hum Mutat 30:
1256-1266.
Heist EK, Ruskin JN (2010). Drug-induced arrhythmia. Circulation 122: 1426-1435.
Hershko A, Ciechanover A (1998). The ubiquitin system. Annu Rev Biochem 67:
425-479.
Herzog RI, Liu C, Waxman SG, Cummins TR (2003). Calmodulin binds to the C
terminus of sodium channels Nav1.4 and Nav1.6 and differentially modulates their
functional properties. J Neurosci 23: 8261-8271.
Hodgson-Zingman DM, Karst ML, Zingman LV, Heublein DM, Darbar D, Herron KJ et
al (2008). Atrial natriuretic peptide frameshift mutation in familial atrial fibrillation. N
Engl J Med 359: 158-165.
109
Hong K, Piper DR, Diaz-Valdecantos A, Brugada J, Oliva A, Burashnikov E et al (2005).
De novo KCNQ1 mutation responsible for atrial fibrillation and short QT syndrome in
utero. Cardiovasc Res 68: 433-440.
Huang H, Zhao J, Barrane FZ, Champagne J, Chahine M (2006). Nav1.5/R1193Q
polymorphism is associated with both long QT and Brugada syndromes. Can J Cardiol
22: 309-313.
Hwang HW, Chen JJ, Lin YJ, Shieh RC, Lee MT, Hung SI et al (2005). R1193Q of
SCN5A, a Brugada and long QT mutation, is a common polymorphism in Han Chinese. J
Med Genet 42: e7.
Isenberg G, Klockner U (1982). Calcium tolerant ventricular myocytes prepared by
preincubation in a "KB medium". Pflugers Arch 395: 6-18.
Isom LL, De Jongh KS, Patton DE, Reber BF, Offord J, Charbonneau H et al (1992).
Primary structure and functional expression of the beta 1 subunit of the rat brain sodium
channel. Science 256: 839-842.
Isom LL, Ragsdale DS, De Jongh KS, Westenbroek RE, Reber BF, Scheuer T et al (1995).
Structure and function of the beta 2 subunit of brain sodium channels, a transmembrane
glycoprotein with a CAM motif. Cell 83: 433-442.
110
Itoh H, Shimizu M, Takata S, Mabuchi H, Imoto K (2005). A novel missense mutation in
the SCN5A gene associated with Brugada syndrome bidirectionally affecting blocking
actions of antiarrhythmic drugs. J Cardiovasc Electrophysiol 16: 686-693.
Jeffs GJ, Meloni BP, Bakker AJ, Knuckey NW (2007). The role of the Na(+)/Ca(2+)
exchanger (NCX) in neurons following ischaemia. J Clin Neurosci 14: 507-514.
Jeong SY, Goto J, Hashida H, Suzuki T, Ogata K, Masuda N et al (2000). Identification
of a novel human voltage-gated sodium channel alpha subunit gene, SCN12A. Biochem
Biophys Res Commun 267: 262-270.
Jespersen T, Gavillet B, van Bemmelen MX, Cordonier S, Thomas MA, Staub O et al
(2006). Cardiac sodium channel Na(v)1.5 interacts with and is regulated by the protein
tyrosine phosphatase PTPH1. Biochem Biophys Res Commun 648: 1455-1462.
Juang JM, Huang SK, Tsai CT, Chiang FT, Lin JL, Lai LP et al (2003). Characteristics of
Chinese patients with symptomatic Brugada syndrome in Taiwan. Cardiology 99:
182-189.
Kane GC, Liu XK, Yamada S, Olson TM, Terzic A (2005). Cardiac KATP channels in
health and disease. J Mol Cell Cardiol 38: 937-943.
111
Kapplinger JD, Tester DJ, Alders M, Benito B, Berthet M, Brugada J et al (2010). An
international compendium of mutations in the SCN5A-encoded cardiac sodium channel
in patients referred for Brugada syndrome genetic testing. Heart Rhythm 7: 33-46.
Kearney JA, Plummer NW, Smith MR, Kapur J, Cummins TR, Waxman SG et al (2001).
A gain-of-function mutation in the sodium channel gene Scn2a results in seizures and
behavioral abnormalities. Neuroscience 102: 307-317.
Kerr NC, Holmes FE, Wynick D (2004). Novel isoforms of the sodium channels Nav1.8
and Nav1.5 are produced by a conserved mechanism in mouse and rat. J Biol Chem 279:
24826-24833.
Kim J, Ghosh S, Liu H, Tateyama M, Kass RS, Pitt GS (2004). Calmodulin mediates
Ca2+ sensitivity of sodium channels. J Biol Chem 279: 45004-45012.
Kiyosue T, Arita M (1989). Late sodium current and its contribution to action potential
configuration in guinea pig ventricular myocytes. Circ Res 64: 389-397.
Klabunde RE (2012). Cardiovascular physiology concepts, 2 edn. Lippincott Williams &
Wilkins: Baltimore, MD
Knollmann BC, Roden DM (2008). A genetic framework for improving arrhythmia
therapy. Nature 451: 929-936.
112
Ko SH, Lenkowski PW, Lee HC, Mounsey JP, Patel MK (2005). Modulation of Na(v)1.5
by beta1-- and beta3-subunit co-expression in mammalian cells. Pflugers Arch 449:
403-412.
Kohrman DC, Smith MR, Goldin AL, Harris J, Meisler MH (1996). A missense mutation
in the sodium channel Scn8a is responsible for cerebellar ataxia in the mouse mutant
jolting. J Neurosci 16: 5993-5999.
Kwon HW, Lee SY, Kwon BS, Kim GB, Bae EJ, Kim WH et al (2012). Long QT
syndrome and dilated cardiomyopathy with SCN5A p.R1193Q polymorphism:
cardioverter-defibrillator implantation at 27 months. Pacing Clin Electrophysiol 35:
e243-246.
Kyndt F, Probst V, Potet F, Demolombe S, Chevallier JC, Baro I et al (2001). Novel
SCN5A mutation leading either to isolated cardiac conduction defect or Brugada
syndrome in a large French family. Circulation 104: 3081-3086.
Lalli MJ, Yong J, Prasad V, Hashimoto K, Plank D, Babu GJ et al (2001). Sarcoplasmic
reticulum Ca(2+) atpase (SERCA) 1a structurally substitutes for SERCA2a in the cardiac
sarcoplasmic reticulum and increases cardiac Ca(2+) handling capacity. Circ Res 89:
160-167.
113
Lederer WJ, Tsien RW (1976). Transient inward current underlying arrhythmogenic
effects of cardiotonic steroids in Purkinje fibres. J Physiol 263: 73-100.
Lei M, Huang CL, Zhang Y (2008). Genetic Na+ channelopathies and sinus node
dysfunction. Prog Biophys Mol Biol 98: 171-178.
Lemaillet G, Walker B, Lambert S (2003). Identification of a conserved ankyrin-binding
motif in the family of sodium channel alpha subunits. J Biol Chem 278: 27333-27339.
Lewartowski B, Zdanowski K (1990). Net Ca2+ influx and sarcoplasmic reticulum Ca2+
uptake in resting single myocytes of the rat heart: comparison with guinea-pig. J Mol Cell
Cardiol 22: 1221-1229.
Li CZ, Wang XD, Wang HW, Bian YT, Liu YM (1997). Four types of late Na channel
current in isolated ventricular myocytes with reference to their contribution to the
lastingness of action potential plateau. Sheng Li Xue Bao 49: 241-248.
Li L, Desantiago J, Chu G, Kranias EG, Bers DM (2000). Phosphorylation of
phospholamban and troponin I in beta-adrenergic-induced acceleration of cardiac
relaxation. Am J Physiol Heart Circ Physiol 278: H769-779.
Liu CJ, Dib-Hajj SD, Renganathan M, Cummins TR, Waxman SG (2003). Modulation of
the cardiac sodium channel Nav1.5 by fibroblast growth factor homologous factor 1B. J
114
Biol Chem 278: 1029-1036.
Liu H, Sun HY, Lau CP, Li GR (2007). Regulation of voltage-gated cardiac sodium
current by epidermal growth factor receptor kinase in guinea pig ventricular myocytes. J
Mol Cell Cardiol 42: 760-768.
Liu YM, DeFelice LJ, Mazzanti M (1992). Na channels that remain open throughout the
cardiac action potential plateau. Biophys J 63: 654-662.
London B, Michalec M, Mehdi H, Zhu X, Kerchner L, Sanyal S et al (2007). Mutation in
glycerol-3-phosphate dehydrogenase 1 like gene (GPD1-L) decreases cardiac Na+
current and causes inherited arrhythmias. Circulation 116: 2260-2268.
Luca F, Alfio Q, Alessandro V (2009). Cardiovascular Mathematics: Modeling and
simulation of the circulatory system. springer.
Lundbaek JA, Birn P, Hansen AJ, Søgaard R, Nielsen C, Girshman J et al (2004).
Regulation of sodium channel function by bilayer elasticity: the importance of
hydrophobic coupling. Effects of Micelle-forming amphiphiles and cholesterol. J Gen
Physiol 123: 599-621.
Lupoglazoff JM, Cheav T, Baroudi G, Berthet M, Denjoy I, Cauchemez B et al (2002).
Homozygous SCN5A mutation in long-QT syndrome with functional two-to-one
115
atrioventricular block. Circ Res 89: E16-21.
Maier SK, Westenbroek RE, Yamanushi TT, Dobrzynski H, Boyett MR, Catterall WA et
al (2003). An unexpected requirement for brain-type sodium channels for control of heart
rate in the mouse sinoatrial node. Proc Natl Acad Sci U S A 100: 3507-3512.
Makielski JC, Ye B, Valdivia CR, Pagel MD, Pu J, Tester DJ et al (2003). A ubiquitous
splice variant and a common polymorphism affect heterologous expression of
recombinant human SCN5A heart sodium channels. Circ Res 93: 821-828.
Makiyama T, Akao M, Tsuji K, Doi T, Ohno S, Takenaka K et al (2005). High risk for
bradyarrhythmic complications in patients with Brugada syndrome caused by SCN5A
gene mutations. J Am Coll Cardiol 46: 2100-2106.
Makiyama T, Akao M, Shizuta S, Doi T, Nishiyama K, Oka Y et al (2008). A novel
SCN5A gain-of-function mutation M1875T associated with familial atrial fibrillation. J
Am Coll Cardiol 52: 1326-1334.
Malhotra JD, Koopmann MC, Kazen-Gillespie KA, Fettman N, Hortsch M, Isom LL
(2002). Structural requirements for interaction of sodium channel beta 1 subunits with
ankyrin. J Biol Chem 277: 26681-26688.
Malhotra JD, Thyagarajan V, Chen C, Isom LL (2004). Tyrosine-phosphorylated and
116
nonphosphorylated sodium channel beta1 subunits are differentially localized in cardiac
myocytes. J Biol Chem 279: 40748-40754.
Marban E, Robinson SW, Wier WG (1986). Mechanisms of arrhythmogenic delayed and
early afterdepolarizations in ferret ventricular muscle. J Clin Invest 78: 1185-1192.
Marfatia KA, Harreman MT, Fanara P, Vertino PM, Corbett AH (2001). Identification and
characterization of the human MOG1 gene. Gene 266: 45-56.
Martin MS, Tang B, Papale LA, Yu FH, Catterall WA, Escayg A (2007). The
voltage-gated sodium channel Scn8a is a genetic modifier of severe myoclonic epilepsy
of infancy. Hum Mol Genet 16: 2892-2899.
Matsusue A, Kashiwagi M, Hara K, Waters B, Sugimura T, Kubo S (2012). An autopsy
case of sudden unexpected nocturnal death syndrome with R1193Q polymorphism in the
SCN5A gene. Leg Med (Tokyo) 14: :317-319.
McNair WP, Ku L, Taylor MR, Fain PR, Dao D, Wolfel E et al (2004). SCN5A mutation
associated with dilated cardiomyopathy, conduction disorder, and arrhythmia. Circulation
110: 2163-2167.
McNaughton PA (1991). Fundamental properties of the Na-Ca exchange. An overview.
Ann N Y Acad Sci 639: 2-9.
117
Meadows LS, Isom LL (2005). Sodium channels as macromolecular complexes:
implications for inherited arrhythmia syndromes. Cardiovasc Res 67: 448-458.
Medeiros-Domingo A, Kaku T, Tester DJ, Iturralde-Torres P, Itty A, Ye B et al (2007).
SCN4B-encoded sodium channel beta4 subunit in congenital long-QT syndrome.
Circulation 116: 134-142.
Meyer-Kleine C, Otto M, Zoll B, Koch MC (1994). Molecular and genetic
characterisation of German families with paramyotonia congenita and demonstration of
founder effect in the Ravensberg families. Human Genetics 93: 707-710.
Mitchell GF, Jeron A, Koren G (1998). Measurement of heart rate and Q-T interval in the
conscious mouse. Am J Physiol 274: H747-751.
Mohler PJ, Schott JJ, Gramolini AO, Dilly KW, Guatimosim S, duBell WH et al (2003).
Ankyrin-B mutation causes type 4 long-QT cardiac arrhythmia and sudden cardiac death.
Nature 421: 634-639.
Mohler PJ, Rivolta I, Napolitano C, LeMaillet G, Lambert S, Priori SG et al (2004).
Nav1.5 E1053K mutation causing Brugada syndrome blocks binding to ankyrin-G and
expression of Nav1.5 on the surface of cardiomyocytes. Proc Natl Acad Sci U S A 101:
17533-17538.
118
Morgan K, Stevens EB, Shah B, Cox PJ, Dixon AK, Lee K et al (2000). beta 3: an
additional auxiliary subunit of the voltage-sensitive sodium channel that modulates
channel gating with distinct kinetics. Proc Natl Acad Sci U S A 97: 2308-2313.
Morrison DK (2009). The 14-3-3 proteins: integrators of diverse signaling cues that
impact cell fate and cancer development. Trends Cell Biol 19: 16-22.
Morrow DA, Scirica BM, Karwatowska-Prokopczuk E, Murphy SA, Budaj A,
Varshavsky S et al (2007). Effects of ranolazine on recurrent cardiovascular events in
patients with non-ST-elevation acute coronary syndromes: the MERLIN-TIMI 36
randomized trial. JAMA 297: 1775-1783.
Mulley JC, Scheffer IE, Petrou S, Dibbens LM, Berkovic SF, Harkin LA (2005). SCN1A
mutations and epilepsy. Hum Mutat 25: 535-542.
Murphy BJ, Rogers J, Perdichizzi AP, Colvin AA, Catterall WA (1996). cAMP-dependent
phosphorylation of two sites in the alpha subunit of the cardiac sodium channel. J Biol
Chem 271: 28837-28843.
Murphy E, Perlman M, London RE, Steenbergen C (1991). Amiloride delays the
ischemia-induced rise in cytosolic free calcium. Circ Res 68: 1250-1258.
119
Navedo MF, Cheng EP, Yuan C, Votaw S, Molkentin JD, Scott JD et al (2010). Increased
coupled gating of L-type Ca2+ channels during hypertension and Timothy syndrome.
Circ Res 106: 748-756.
Neyroud N, Tesson F, Denjoy I, Leibovici M, Donger C, Barhanin J et al (1997). A novel
mutation in the potassium channel gene KVLQT1 causes the Jervell and Lange-Nielsen
cardioauditory syndrome. Nat Genet 15: 186-189.
Nguyen SV, Claycomb WC (1999). Hypoxia regulates the expression of the
adrenomedullin and HIF-1 genes in cultured HL-1 cardiomyocytes. Biochem Biophys Res
Commun 265: 382-386.
Nilsen KB, Nicholas AK, Woods CG, Mellgren SI, Nebuchennykh M, Aasly J (2009).
Two novel SCN9A mutations causing insensitivity to pain. Pain 143: 155-158.
Oberti C, Wang L, Li L, Dong J, Rao S, Du W et al (2004). Genome-wide linkage scan
identifies a novel genetic locus on chromosome 5p13 for neonatal atrial fibrillation
associated with sudden death and variable cardiomyopathy. Circulation 110: 3753-3759.
Oki M, Nishimoto T (1998). A protein required for nuclear-protein import, Mog1p,
directly interacts with GTP-Gsp1p, the Saccharomyces cerevisiae ran homologue. Proc
Natl Acad Sci U S A 95: 15388-15393.
120
Olsen SK, Garbi M, Zampieri N, Eliseenkova AV, Ornitz DM, Goldfarb M et al (2003).
Fibroblast growth factor (FGF) homologous factors share structural but not functional
homology with FGFs. J Biol Chem 278: 34226-34236.
Olson TM, Keating MT (1996). Mapping a cardiomyopathy locus to chromosome
3p22-p25. J Clin Invest 97: 528-532.
Olson TM, Michels VV, Ballew JD, Reyna SP, Karst ML, Herron KJ et al (2005).
Sodium channel mutations and susceptibility to heart failure and atrial fibrillation. JAMA
293: 447-454.
Olson TM, Alekseev AE, Liu XK, Park S, Zingman LV, Bienengraeber M et al (2006).
Kv1.5 channelopathy due to KCNA5 loss-of-function mutation causes human atrial
fibrillation. Hum Mol Genet 15: 2185-2191.
Otway R, Vandenberg JI, Guo G, Varghese A, Castro ML, Liu J et al (2007).
Stretch-sensitive KCNQ1 mutation A link between genetic and environmental factors in
the pathogenesis of atrial fibrillation. J Am Coll Cardiol 49: 578-586.
Palade GE (1953). An electron microscope study of the mitochondrial structure. J
Histochem Cytochem 1: 188-211.
Papadatos GA, Wallerstein PM, Head CE, Ratcliff R, Brady PA, Benndorf K et al (2002).
121
Slowed conduction and ventricular tachycardia after targeted disruption of the cardiac
sodium channel gene Scn5a. Proc Natl Acad Sci U S A 99: 6210-6215.
Patino GA, Isom LL (2010). Electrophysiology and beyond: multiple roles of Na+
channel β subunits in development and disease. Neurosci Lett 486: 53-59.
Picht E, DeSantiago J, Blatter LA, Bers DM (2006). Cardiac alternans do not rely on
diastolic sarcoplasmic reticulum calcium content fluctuations. Circ Res 99: 740-748.
Picht E, DeSantiago J, Huke S, Kaetzel MA, Dedman JR, Bers DM (2007). CaMKII
inhibition targeted to the sarcoplasmic reticulum inhibits frequency-dependent
acceleration of relaxation and Ca2+ current facilitation. J Mol Cell Cardiol 42: 196-205.
Pieske B, Maier LS, Piacentino V, 3rd, Weisser J, Hasenfuss G, Houser S (2002). Rate
dependence of [Na+]i and contractility in nonfailing and failing human myocardium.
Circulation 106: 447-453.
Piper HM (2000). The calcium paradox revisited: an artefact of great heuristic value.
Cardiovasc Res 45: 123-127.
Plassart E, Reboul J, Rime C, Recan D, Millasseau P, Eymard B et al (1994). Mutations
in the muscle sodium channel gene (SCN4A) in 13 French families with hyperkalemic
periodic paralysis and paramyotonia congenita: phenotype to genotype correlations and
122
demonstration of the predominance of two mutations. Eur J Hum Genet 2: 110-124.
Plummer NW, Galt J, Jones JM, Burgess DL, Sprunger LK, Kohrman DC et al (1998).
Exon organization, coding sequence, physical mapping, and polymorphic intragenic
markers for the human neuronal sodium channel gene SCN8A. Genomics 54: 287-296.
Plummer NW, Meisler MH (1999). Evolution and diversity of mammalian sodium
channel genes. Genomics 57: 323-331.
Pogwizd SM, Schlotthauer K, Li L, Yuan W, Bers DM (2001). Arrhythmogenesis and
contractile dysfunction in heart failure: Roles of sodium-calcium exchange, inward
rectifier potassium current, and residual beta-adrenergic responsiveness. Circ Res 88:
1159-1167.
Pogwizd SM, Bers DM (2004). Cellular basis of triggered arrhythmias in heart failure.
Trends Cardiovasc Med 14: 61-66.
Potet F, Mabo P, Le Coq G, Probst V, Schott JJ, Airaud F et al (2003). Novel brugada
SCN5A mutation leading to ST segment elevation in the inferior or the right precordial
leads. J Cardiovasc Electrophysiol 14: 200-203.
Potts JF, Regan MR, Rochelle JM, Seldin MF, Agnew WS (1993). A glial-specific
voltage-sensitive Na channel gene maps close to clustered genes for neuronal isoforms on
123
mouse chromosome 2. Biochem Biophys Res Commun 197: 100-104.
Qiu X, Liu W, Hu D, Sun Y, Li L, Li C (2009). Patient with obstructive sleep
apnea-hypopnea syndrome and SCN5A mutation (R1193Q polymorphism) associated
with Brugada type 2 electrocardiographic pattern. J Electrocardiol 42: 250-253.
Qu Y, Rogers J, Tanada T, Scheuer T, Catterall WA (1994). Modulation of cardiac Na+
channels expressed in a mammalian cell line and in ventricular myocytes by protein
kinase C. Proc Natl Acad Sci U S A 91: 3289-3293.
Qu Y, Isom LL, Westenbroek RE, Rogers JC, Tanada TN, McCormick KA et al (1995).
Modulation of cardiac Na+ channel expression in Xenopus oocytes by beta 1 subunits. J
Biol Chem 270: 25696-25701.
Ragsdale DS, McPhee JC, Scheuer T, Catterall WA (1994). Molecular determinants of
state-dependent block of Na+ channels by local anesthetics. Science 265: 1724-1728.
Ravn LS, Aizawa Y, Pollevick GD, Hofman-Bang J, Cordeiro JM, Dixen U et al (2008).
Gain of function in IKs secondary to a mutation in KCNE5 associated with atrial
fibrillation. Heart Rhythm 5: 427-435.
Reddy BM, Weintraub HS, Schwartzbard AZ (2010). Ranolazine: a new approach to
treating an old problem. Tex Heart Inst J 37: 641-647.
124
Remme CA, Wilde AA, Bezzina CR (2008). Cardiac sodium channel overlap syndromes:
different faces of SCN5A mutations. Trends Cardiovasc Med 18: 78-87.
Ritter M, Su Z, Xu S, Shelby J, Barry WH (2000). Cardiac unloading alters contractility
and calcium homeostasis in ventricular myocytes. J Mol Cell Cardiol 32: 577-584.
Roden DM, Woosley RL, Primm RK (1986). Incidence and clinical features of the
quinidine-associated long QT syndrome: implications for patient care. Am Heart J 111:
1088-1093.
Roden DM (1994). Risks and benefits of antiarrhythmic therapy. N Engl J Med 331:
785-791.
Rook MB, Alshinawi CB, Groenewegen WA, van Gelder IC, van Ginneken AC, Jongsma
HJ et al (1999). Human SCN5A gene mutations alter cardiac sodium channel kinetics and
are associated with the Brugada syndrome. Cardiovasc Res 44: 507-517.
Rossenbacker T, Carroll SJ, Liu H, Kuipéri C, de Ravel TJ, Devriendt K et al (2004).
Novel pore mutation in SCN5A manifests as a spectrum of phenotypes ranging from
atrial flutter, conduction disease, and Brugada syndrome to sudden cardiac death. Heart
Rhythm 1: 610-615.
125
Rota M, Boni A, Urbanek K, Padin-Iruegas ME, Kajstura TJ, Fiore G et al (2005).
Nuclear targeting of Akt enhances ventricular function and myocyte contractility. Circ
Res 97: 1332-1341.
Rotin D, Kumar S (2009). Physiological functions of the HECT family of ubiquitin
ligases. Nat Rev Mol Cell Biol 10: 398-409.
Rougier JS, van Bemmelen MX, Bruce MC, Jespersen T, Gavillet B, Apothéloz F et al
(2005). Molecular determinants of voltage-gated sodium channel regulation by the
Nedd4/Nedd4-like proteins. Am J Physiol Cell Physiol 288: C692-701.
Royer A, van Veen TA, Le Bouter S, Marionneau C, Griol-Charhbili V, Léoni AL et al
(2005). Mouse model of SCN5A-linked hereditary Lenègre's disease: age-related
conduction slowing and myocardial fibrosis. Circulation 111: 1738-1746.
Ruan Y, Liu N, Bloise R, Napolitano C, Priori SG (2007). Gating properties of SCN5A
mutations and the response to mexiletine in long-QT syndrome type 3 patients.
Circulation 116: 1137-1144.
Ruan Y, Liu N, Priori SG (2009). Sodium channel mutations and arrhythmias. Nat Rev
Cardiol 6: 337-348.
Saimi Y, Kung C (2002). Calmodulin as an ion channel subunit. Annu Rev Physiol 64:
126
289-311.
Saint DA (2008). The cardiac persistent sodium current: an appealing therapeutic target.
Br J Pharmacol 153: 1133-1142.
Saitoh M, Shinohara M, Hoshino H, Kubota M, Amemiya K, Takanashi JL et al (2012).
Mutations of the SCN1A gene in acute encephalopathy. Epilepsia 53: 558-564.
Sandow A (1952). Excitation-contraction coupling in muscular response. Yale J Biol Med
25: 176-201.
Schmitt JP, Kamisago M, Asahi M, Li GH, Ahmad F, Mende U et al (2003). Dilated
cardiomyopathy and heart failure caused by a mutation in phospholamban. Science 299:
1410-1413.
Schott JJ, Alshinawi C, Kyndt F, Probst V, Hoorntje TM, Hulsbeek M et al (1999).
Cardiac conduction defects associate with mutations in SCN5A. Nat Genet 23: 20-21.
Schroeter A, Walzik S, Blechschmidt S, Haufe V, Benndorf K, Zimmer T (2010).
Structure and function of splice variants of the cardiac voltage-gated sodium channel
Na(v)1.5. J Mol Cell Cardiol 49: 16-24.
Schulze-Bahr E, Eckardt L, Breithardt G, Seidl K, Wichter T, Wolpert C et al (2003).
127
Sodium channel gene (SCN5A) mutations in 44 index patients with Brugada syndrome:
different incidences in familial and sporadic disease. Hum Mutat 21: 651-652.
Sedej S, Heinzel FR, Walther S, Dybkova N, Wakula P, Groborz J et al (2010).
Na+-dependent SR Ca2+ overload induces arrhythmogenic events in mouse
cardiomyocytes with a human CPVT mutation. Cardiovasc Res 87: 50-59.
Selzer A, Wray HW (1964). Qqinidine syncope. paroxysmal ventricular fibrillation
occurring during treatment of chronic atrial arrhythmias. Circulation 30: 17-26.
Shah VN, Wingo TL, Weiss KL, Williams CK, Balser JR, Chazin WJ (2006).
Calcium-dependent regulation of the voltage-gated sodium channel hH1: intrinsic and
extrinsic sensors use a common molecular switch. Proc Natl Acad Sci U S A 103:
3592-3597.
Sheets MF, Fozzard HA, Lipkind GM, Hanck DA (2010). Sodium channel molecular
conformations and antiarrhythmic drug affinity. Trends Cardiovasc Med 20: 16-21.
Shim SH, Ito M, Maher T, Milunsky A (2005). Gene sequencing in neonates and infants
with the long QT syndrome. Genet Test 9: 281-284.
Shin DJ, Kim E, Park SB, Jang WC, Bae Y, Han J et al (2007). A novel mutation in the
SCN5A gene is associated with Brugada syndrome. Life Sci 80: 719-727.
128
Sicouri S, Antzelevitch D, Heilmann C, Antzelevitch C (1997). Effects of sodium channel
block with mexiletine to reverse action potential prolongation in in vitro models of the
long term QT syndrome. J Cardiovasc Electrophysiol 8: 1280-1290.
Skinner JR, Chung SK, Montgomery D, McCulley CH, Crawford J, French J et al (2005).
Near-miss SIDS due to Brugada syndrome. Arch Dis Child 90: 528-529.
Smits JP, Koopmann TT, Wilders R, Veldkamp MW, Opthof T, Bhuiyan ZA et al (2005).
A mutation in the human cardiac sodium channel (E161K) contributes to sick sinus
syndrome, conduction disease and Brugada syndrome in two families. J Mol Cell Cardiol
38: 969-981.
Song KS, Scherer PE, Tang Z, Okamoto T, Li S, Chafel M et al (1996). Expression of
caveolin-3 in skeletal, cardiac, and smooth muscle cells. Caveolin-3 is a component of
the sarcolemma and co-fractionates with dystrophin and dystrophin-associated
glycoproteins. J Biol Chem 271: 160-165.
Song Y, Shryock JC, Wagner S, Maier LS, Belardinelli L (2006). Blocking late sodium
current reduces hydrogen peroxide-induced arrhythmogenic activity and contractile
dysfunction. J Pharmacol Exp Ther 318: 214-222.
Splawski I, Timothy KW, Vincent GM, Atkinson DL, Keating MT (1997a). Molecular
129
basis of the long-QT syndrome associated with deafness. N Engl J Med 336: 1562-1567.
Splawski I, Tristani-Firouzi M, Lehmann MH, Sanguinetti MC, Keating MT (1997b).
Mutations in the hminK gene cause long QT syndrome and suppress IKs function. Nat
Genet 17: 338-340.
Splawski I, Shen J, Timothy KW, Lehmann MH, Priori S, Robinson JL et al (2000).
Spectrum of mutations in long-QT syndrome genes. KVLQT1, HERG, SCN5A, KCNE1,
and KCNE2. Circulation 102: 1178-1185.
Staub O, Rotin D (2006). Role of ubiquitylation in cellular membrane transport. Physiol
Rev 86: 669-677.
Stone PH, Gratsiansky NA, Blokhin A, Huang IZ, Meng L (2006). Antianginal efficacy
of ranolazine when added to treatment with amlodipine: the ERICA (Efficacy of
Ranolazine in Chronic Angina) trial. J Am Coll Cardiol 48: 566-575.
Sun A, Xu L, Wang S, Wang K, Huang W, Wang Y et al (2008). SCN5A R1193Q
polymorphism associated with progressive cardiac conduction defects and long QT
syndrome in a Chinese family. J Med Genet 45: 127-128.
Takehara N, Makita N, Kawabe J, Sato N, Kawamura Y, Kitabatake A et al (2004). A
cardiac sodium channel mutation identified in Brugada syndrome associated with atrial
130
standstill. J Intern Med 255: 137-142.
Tan HL, Bink-Boelkens MT, Bezzina CR, Viswanathan PC, Beaufort-Krol GC, van
Tintelen PJ et al (2001). A sodium-channel mutation causes isolated cardiac conduction
disease. Nature 409: 1043-1047.
Tan HL, Kupershmidt S, Zhang R, Stepanovic S, Roden DM, Wilde AA et al (2002). A
calcium sensor in the sodium channel modulates cardiac excitability. Nature 415:
442-447.
Tang Z, Scherer PE, Okamoto T, Song K, Chu C, Kohtz DS et al (1996). Molecular
cloning of caveolin-3, a novel member of the caveolin gene family expressed
predominantly in muscle. J Biol Chem 274: 2255-2261.
Tian XL, Yong SL, Wan X, Wu L, Chung MK, Tchou PJ et al (2004). Mechanisms by
which SCN5A mutation N1325S causes cardiac arrhythmias and sudden death in vivo.
Cardiovasc Res 61: 256-267.
Tian XL, Wang QK (2006). Generation of transgenic mice for cardiovascular research.
Methods Mol Med 129: 68-81.
Tian XL, Cheng Y, Zhang T, Liao ML, Yong SL, Wang QK (2007). Optical mapping of
ventricular arrhythmias in LQTS mice with SCN5A mutation N1325S. Biochem Biophys
131
Res Commun 352: 879-883.
Tristani-Firouzi M, Jensen JL, Donaldson MR, Sansone V, Meola G, Hahn A et al (2002).
Functional and clinical characterization of KCNJ2 mutations associated with LQT7
(Andersen syndrome). J Clin Invest 110: 381-388.
Trudeau MM, Dalton JC, Day JW, Ranum LP, Meisler MH (2006). Heterozygosity for a
protein truncation mutation of sodium channel SCN8A in a patient with cerebellar
atrophy, ataxia, and mental retardation. J Med Genet 43: 527-530.
Ueda K, Valdivia C, Medeiros-Domingo A, Tester DJ, Vatta M, Farrugia G et al (2008).
Syntrophin mutation associated with long QT syndrome through activation of the
nNOS-SCN5A macromolecular complex. Proc Natl Acad Sci U S A 105: 9355-9360.
Undrovinas AI, Shander GS, Makielski JC (1995). Cytoskeleton modulates gating of
voltage-dependent sodium channel in heart. Am J Physiol 269: H203-214.
Valdivia CR, Ackerman MJ, Tester DJ, Wada T, McCormack J, Ye B et al (2002). A novel
SCN5A arrhythmia mutation, M1766L, with expression defect rescued by mexiletine.
Cardiovasc Res 55: 279-289.
Van Norstrand DW, Valdivia CR, Tester DJ, Ueda K, London B, Makielski JC et al
(2007). Molecular and functional characterization of novel glycerol-3-phosphate
132
dehydrogenase 1 like gene (GPD1-L) mutations in sudden infant death syndrome.
Circulation 116: 2253-2259.
Veldkamp MW, Wilders R, Baartscheer A, Zegers JG, Bezzina CR, Wilde AA (2003).
Contribution of sodium channel mutations to bradycardia and sinus node dysfunction in
LQT3 families. Circ Res 92: 976-913.
Vincent GM (2003). Romano-Ward Syndrome. Unviersity of Washington: Seattle (WA).
Wagner S, Dybkova N, Rasenack EC, Jacobshagen C, Fabritz L, Kirchhof P et al (2006).
Ca2+/calmodulin-dependent protein kinase II regulates cardiac Na+ channels. J Clin
Invest 116: 3127-3138.
Wallace RH, Scheffer IE, Barnett S, Richards M, Dibbens L, Desai RR et al (2001).
Neuronal sodium-channel alpha1-subunit mutations in generalized epilepsy with febrile
seizures plus. Am J Hum Genet 68: 859-865.
Wan X, Chen S, Sadeghpour A, Wang Q, Kirsch GE (2001). Accelerated inactivation in a
mutant Na(+) channel associated with idiopathic ventricular fibrillation. Am J Physiol
Heart Circ Physiol 280: H354-360.
Wang DW, Desai RR, Crotti L, Arnestad M, Insolia R, Pedrazzini M et al (2007). Cardiac
sodium channel dysfunction in sudden infant death syndrome. Circulation 115: 368-376.
133
Wang L, Lopaschuk GD, Clanachan AS (2008). H(2)O(2)-induced left ventricular
dysfunction in isolated working rat hearts is independent of calcium accumulation. J Mol
Cell Cardiol 45: 787-795.
Wang P, Yang Q, Wu X, Yang Y, Shi L, Wang C et al (2010). Functional
dominant-negative mutation of sodium channel subunit gene SCN3B associated with
atrial fibrillation in a Chinese GeneID population. Biochem Biophys Res Commun 398:
98-104.
Wang Q, Shen J, Li Z, Timothy K, Vincent GM, Priori SG et al (1995a). Cardiac sodium
channel mutations in patients with long QT syndrome, an inherited cardiac arrhythmia.
Hum Mol Genet 4: 1603-1607.
Wang Q, Shen J, Splawski I, Atkinson D, Li Z, Robinson JL et al (1995b). SCN5A
mutations associated with an inherited cardiac arrhythmia, long QT syndrome. Cell 80:
805-811.
Wang Q, Curran ME, Splawski I, Burn TC, Millholland JM, VanRaay TJ et al (1996a).
Positional cloning of a novel potassium channel gene: KVLQT1 mutations cause cardiac
arrhythmias. Nat Genet 12: 17-23.
Wang Q, Li Z, Shen J, Keating MT (1996b). Genomic organization of the human SCN5A
134
gene encoding the cardiac sodium channel. Genomics 34: 9-16.
Wang Q, Chen S, Chen Q, Wan X, Shen J, Hoeltge GA et al (2004). The common
SCN5A mutation R1193Q causes LQTS-type electrophysiological alterations of the
cardiac sodium channel. J Med Genet 41: e66.
Wang TJ, Larson MG, Levy D, Vasan RS, Leip EP, Wolf PA et al (2003). Temporal
relations of atrial fibrillation and congestive heart failure and their joint influence on
mortality: the Framingham Heart Study. Circulation 107: 2920-2925.
Wang Z, Fermini B, Nattel S (1995c). Effects of flecainide, quinidine, and
4-aminopyridine on transient outward and ultrarapid delayed rectifier currents in human
atrial myocytes. J Pharmacol Exp Ther 272: 184-196.
Watanabe H, Darbar D, Kaiser DW, Jiramongkolchai K, Chopra S, Donahue BS et al
(2009). Mutations in sodium channel beta1- and beta2-subunits associated with atrial
fibrillation. Circ Arrhythm Electrophysiol 2: 268-275.
Wattanasirichaigoon D, Vesely MR, Duggal P, Levine JC, Blume ED, Wolff GS et al
(1999). Sodium channel abnormalities are infrequent in patients with long QT syndrome:
identification of two novel SCN5A mutations. Am J Med Genet 86: 470-476.
Wei J, Wang DW, Alings M, Fish F, Wathen M, Roden DM et al (1999). Congenital
135
long-QT syndrome caused by a novel mutation in a conserved acidic domain of the
cardiac Na+ channel. Circulation 99: 3165-3171.
Weiss R, Barmada MM, Nguyen T, Seibel JS, Cavlovich D, Kornblit CA et al (2002).
Clinical and molecular heterogeneity in the Brugada syndrome: a novel gene locus on
chromosome 3. Circulation 105: 707-713.
White SM, Constantin P, Claycomb WC (2004). Cardiac physiology at the cellular level:
use of cultured HL-1 cardiomyocytes for studies of cardiac muscle cell structure and
function. Am J Physiol Heart Circ Physiol 286: H823-829.
Wier WG (1990). Cytoplasmic [Ca2+] in mammalian ventricle: dynamic control by
cellular processes. Annu Rev Physiol 52: 467-485.
Williams IA, Xiao XH, Ju YK, Allen DG (2007). The rise of [Na(+)] (i) during ischemia
and reperfusion in the rat heart-underlying mechanisms. Pflugers Arch 454: 903-912.
Williams TM, Lisanti MP (2004). The Caveolin genes: from cell biology to medicine.
Ann Med 36: 584-595.
Wingo TL, Shah VN, Anderson ME, Lybrand TP, Chazin WJ, Balser JR (2004). An
EF-hand in the sodium channel couples intracellular calcium to cardiac excitability. Nat
Struct Mol Biol 11: 219-225.
136
Woosley RL, Chen Y, Freiman JP, Gillis RA (1993). Mechanism of the cardiotoxic
actions of terfenadine. JAMA 269: 1532-1536.
Wu L, Yong SL, Fan C, Ni Y, Yoo S, Zhang T et al (2008). Identification of a new
co-factor, MOG1, required for the full function of cardiac sodium channel Nav 1.5. J Biol
Chem 283: 6968-6978.
Wwidmann S (1955). Effects of calcium ions and local anesthetics on electrical
properties of Purkinje fibres. J Physiol 129: 568-582.
Yang Y, Xia M, Jin Q, Bendahhou S, Shi J, Chen Y et al (2004). Identification of a
KCNE2 gain-of-function mutation in patients with familial atrial fibrillation. Am J Hum
Genet 75: 899-905.
Yang Y, Yang Y, Liang B, Liu J, Li J, Grunnet M et al (2010). Identification of a Kir3.4
mutation in congenital long QT syndrome. Am J Hum Genet 86: 872-880.
Yarbrough TL, Lu T, Lee HC, Shibata EF (2002). Localization of cardiac sodium
channels in caveolin-rich membrane domains: regulation of sodium current amplitude.
Circ Res 90: 443-449.
Yong SL, Ni Y, Zhang T, Tester DJ, Ackerman MJ, Wang QK (2007). Characterization of
137
the cardiac sodium channel SCN5A mutation, N1325S, in single murine ventricular
myocytes. Biochem Biophys Res Commun 352: 378-393.
Yu FH, Westenbroek RE, Silos-Santiago I, McCormick KA, Lawson D, Ge P et al (2003).
Sodium channel beta4, a new disulfide-linked auxiliary subunit with similarity to beta2. J
Neurosci 23: 7577-7585.
Zaza A, Belardinelli L, Shryock JC (2008). Pathophysiology and pharmacology of the
cardiac "late sodium current.". Pharmacol Ther 119: 326-339.
Zhang KX, Zhu DL, He X, Zhang Y, Zhang H, Zhao R et al (2003). [Association of
single nucleotide polymorphism in human SCN7A gene with essential hypertension in
Chinese]. Zhonghua Yi Xue Yi Chuan Xue Za Zhi 20: 463-467.
Zhang T, Yong S, Drink J, Popvic Z, Wang Q (2006). Late sodium currents generated by
mutation N1325S in sodium channel gene SCN5A cause heart failure. Circulation 114:
II-65.
Zhang T, Yong SL, Tian XL, Wang QK (2007). Cardiac-specific overexpression of
SCN5A gene leads to shorter P wave duration and PR interval in transgenic mice.
Biochem Biophys Res Commun 355: 444-450.
Zhang T, Yong SL, Drinko JK, Popović ZB, Shryock JC, Belardinelli L et al (2011).
138
LQTS mutation N1325S in cardiac sodium channel gene SCN5A causes cardiomyocyte
apoptosis, cardiac fibrosis and contractile dysfunction in mice. Int J Cardiol 147:
239-245.
Zhang X, Chen S, Yoo S, Chakrabarti S, Zhang T, Ke T et al (2008). Mutation in nuclear
pore component NUP155 leads to atrial fibrillation and early sudden cardiac death. Cell
135: 1017-1027.
Zhou J, Yi J, Hu N, George AL, Jr., Murray KT (2000). Activation of protein kinase A
modulates trafficking of the human cardiac sodium channel in Xenopus oocytes. Circ Res
87: 33-38.
139